You are on page 1of 30

Raman Spectroscopy of 7 Operando and Time-Resolved,

Space-resolved, and Variable Temperature-


Catalysts Programmed Spectroscopy 16
7.1 Raman Spectra during catalyst synthesis 16
8 Raman Spectra During Treatments 16
Miguel A. Bañares
8.1 Temperature-programmed Reduction-/
Catalytic Spectroscopy Laboratory, Insituto de Temperature-programmed Oxidation-
Catálisis, CSIC, Madrid, Spain Raman 16
8.2 Temperature-programmed Desorption
Israel E. Wachs and Temperature-programmed Surface
Operando Molecular Spectroscopy & Catalysis Reaction-Raman 17
Laboratory, Chemical Engineering Department, 9 Operando Raman Studies and Raman
Lehigh University, Bethlehem, PA, USA Spectra During Reaction with Simultaneous
Activity/Selectivity Determination 17
9.1 VSbO4 Transformations During
1 Introduction 2 Propane Ammoxidation Reaction:
2 Raman Spectroscopy 3 A Temperature-resolved Operando
2.1 Raman Effect 3 Raman-Gas Chromatography Study 17
2.2 Raman Intensity versus Excitation 9.2 Metal Molybdate Bulk-phase
Wavelength 4 Transitions During Reaction, a
2.3 Benefits and Limitations of the Time-resolved Operando Resonance
Different Excitation Lines in Raman Raman-Gas Chromatography Study 18
Spectroscopy 5 9.3 Raman Spectra of Redisper-
2.4 Raman Benefits versus Infrared 6 sion/Regeneration During Reaction 19
2.5 Raman Limitations versus Infrared 7 10 Chemical Image Microscopy 19
2.6 ‘‘Mass Effect’’, Isotopic Substitutions 8 10.1 Raman Microspectroscopy, Scanning
2.7 The Diatomic Approximation 8 Near-field Optical Microscopy and
Tip-enhanced Raman Spectroscopy 20
3 Raman System 9
10.2 Operando Time- and Space-resolved
3.1 Excitation and Laser-line Rejection in
Raman 22
Spectrometer 9
11 Multitechniques 23
3.2 Detectors 9
Abbreviations and Acronyms 23
3.3 Calibration in Intensity and Wavelength 9
Related Articles 23
3.4 Optical Sampling System 9
References 24
4 Raman Sampling for Characterization of
Samples in Reactive Environments 10 Further Reading 29
5 Equipment for Raman Spectroscopy of
Catalysts in Reactive Atmospheres 10
5.1 Reactor Cells for Raman Spectroscopy
in the Liquid Phase 10 This chapter focuses on the application of Raman spec-
5.2 Cells for Raman Spectroscopy of Solids troscopy for characterizing catalytic materials and for the
in Reactive Environment 11 investigation of the bulk and surface chemistry occur-
6 Raman in Catalysis, A Journey that ring during catalyst preparation and operation; a general
Started in the 1970s 13 theoretical background and description of equipment
for Raman experimentation is also presented. Over the
6.1 Early Vacuum, Chemisorption, and
past decades, Raman spectroscopy has increasingly been
Hydration/Dehydration Experiments 13
applied for characterization of all types of catalytic mate-
6.2 Hydrated Supported Metal Oxides 13 rials: bulk and supported metals, bulk mixed metal oxides,
6.3 Dehydrated Supported Metal Oxide supported metal oxides, bulk and supported metal sulfides,
Catalysts 14 zeolites and molecular sieves, heteropolyoxo anions, and
6.4 Hydrated and Dehydrated Bulk Metal clays. Raman is also used to study chemisorption and,
Oxide Catalysts 15 in recent years, Raman has been increasingly applied

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
2 RAMAN SPECTROSCOPY

under in situ conditions. There is an increasingly impor- Table 1 Spectroscopic techniques suitable for in situ
tant strong driving force leading research toward the study characterization
of catalysts under reaction conditions. The last few years Technique Outline
exhibit an increasing number of Raman studies during rele-
UV–Vis Detects electronic transitions and weak
vant reaction conditions (operando methodology). This bands for reduced states at short
momentum in reaction and operando Raman spectroscopy acquisition times
advances a new revolution in catalysis science that is poised ESR/EPR Detects oxidation states of paramagnetic
to afford accurate molecular understanding of catalytic ions at low temperatures and requires
structure–activity/selectivity relationships. long acquisition times
NMR Analyzes local structure, but has long
acquisition times, limited number of
elements
XAS Analyzes local structure and oxidation
1 INTRODUCTION state
Infrared, Raman Provides molecular structural information
Catalysis plays a key role in the chemical industry since at short acquisition times
Mössbauer Analyzes local environment, but is limited
most products are produced via a catalytic process or to few elements like Fe, Co, or Sb
from an intermediate produced via a catalytic process.
Therefore, understanding catalytic phenomenon has
significant fundamental and applied relevance. Spectro- pressures allows for the presence of weakly adsorbed
scopic techniques are extremely powerful for catalyst intermediates in significant concentrations, which would
characterization because they can provide fundamental not be observed under vacuum.(11 – 13) These ‘‘gaps’’ in
information about catalytic structures and surface reac- materials and pressure have to be bridged to estab-
tion intermediates under controlled environments(1 – 3) lish quantitative structure–activity relationships.(2,17 – 19)
(Table 1). The techniques listed in Table 1 provide funda- Raman spectroscopy is one of the most useful tech-
mental information about different aspects of catalysts. niques for bridging these ‘‘gaps’’ and is the focus of this
UV–vis (ultraviolet–visible) provides details about the article. Significant changes in catalyst structures result
oxidation state of cations, their environment and nucle- from changes in operating conditions and are reflected
arity, based on electronic transitions. ESR (electron in the corresponding catalytic performance. The simul-
spin resonance) detects oxidation states of paramagnetic taneous determination of catalyst structure and activity
elements and affords indication about their coordina- or selectivity is needed to establish structure–activity or
tion environment. NMR (nuclear magnetic resonance) structure–selectivity relationships, which can be the basis
is sensitive to local structures of several nuclei. X-ray for catalyst improvement and development.(1,2,10,20,21)
absorption spectroscopy informs on the local geometry A number of monographs and review articles discuss
of the catalytic active sites and some surface reaction Raman spectroscopy in heterogeneous catalysis.(20,22 – 65)
intermediates. Vibrational spectroscopies, like infrared The general physics of Raman spectroscopy is introduced
(IR) or Raman, provide molecular structural informa- only briefly here. The advantages and disadvantages of
tion. Mössbauer is very informative, but limited to few Raman spectroscopy (e.g. quantification problems) are
elements, like Fe, Co, or Sb. The combination of funda- also discussed. The need for characterization of catalysts
mental structural information with in situ capabilities trig- during operation has been emphasized and demonstrated
gers understanding in catalysis science, which allows for by many authors.(2,3,10,22,23,30,39,40,43,44,46,58,63 – 74) Raman
the development of a molecular-level understanding of spectroscopy has emerged as one of the most powerful
the structure–activity/selectivity relationship of catalytic tools for characterization of working catalysts. Raman
reactions. experiments can be carried out at temperatures above
Spectroscopic techniques for characterization of cata- 1000 ° C and at elevated pressures. For materials with high
lysts in the working state are informative, because they Raman cross sections, it is possible to achieve subsecond
provide fundamental insights about catalyst structures time resolutions. Thus, time-resolved transient tempera-
and surface reaction intermediates. These studies on ture or pressure response experiments can be carried out.
model surfaces(1 – 10) permit major advances in catalysis Furthermore, quartz fiber optics allows easy spectroscopic
science, and can be the basis for the design or discovery access to catalytic reactors and vessels.
of new catalysts. Numerous spectroscopic techniques are Raman spectroscopy has been used to examine essen-
applied with model catalysts such as single crystals or well- tially all types of catalytic materials: bulk and supported
defined clusters and under vacuum conditions. Surface metals, bulk mixed metal oxides, supported metal oxides,
structures in polycrystalline catalysts, however, are far bulk and supported metal sulfides, zeolites and molecular
more complex(2,11 – 16) and operation at relatively high sieves, heteropolyoxo anions, and clays.(22,62,75) Over the

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 3

700 Number of publications

600 Operando
Situ
500 Raman

400

300

200

100

0
1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015

Figure 1 Publications per year about Raman spectroscopy in catalysis. Source, Scopus, terms of search (Red) Raman + Catalyst∗ ,
(Green) Raman + Catalyst∗ + situ, (Blue), Raman + catalyst∗ + operando.

past three to four decades, Raman spectroscopy has scattering affords information about the structure and
increasingly been applied for catalyst characterization. properties of molecules from their vibrational transitions.
Figure 1 shows the number of publications about catalysts However, the fundamentals of Raman scattering are
that have used Raman spectroscopy for characteriza- different and complement those of IR absorption. The
tion. This illustration confirms earlier trends reported by IR absorption is linked to a direct resonance between
Wachs in 2001.(75) Most of the Raman work have been the frequency of the IR radiation and the vibrational
on bulk and supported metal oxide catalysts because of frequency of a specific vibration mode. Transitions that
the excellent Raman signals from these type of metal involve changes in the dipole moment of the molecule
oxides.(75) Earlier in situ Raman works can be found during vibration are essentially active in IR. Thus, IR
in the papers published in the late 1970s.(32,34 – 36,50,76) absorption is a one-photon event. The Raman transitions
These papers dealt mainly with chemisorption. In recent among energy levels also correspond with the transitions
years, Raman has been increasingly employed in the between vibrational states, but Raman scattering is a
study of other catalytic materials: molecular sieves, two-photon event. In this case, the property involved is
metals, and heteropolycompounds. Raman studies of the change in the polarizability of the molecule during
bulk and supported metal sulfides, clays, and hydrotal- vibration. The light scattered consists of both elastic
cites, however, are still rather low. The total number radiation (Rayleigh scattering) and inelastic scattering
of Raman studies has been increasing during the last (Raman scattering). The Raman photons are shifted in
few decades; among them only a small fraction is run frequency (energy) from the frequency of the incident
under in situ conditions. There is an increasingly impor- radiation, the shift corresponds to the vibrational energy
tant strong driving force leading research toward the that is gained or lost in the molecule.
study of catalysts under reaction conditions. The last few An energy-level diagram illustrates IR absorption and
years exhibit an increasing number of Raman studies Raman scattering (Figure 2). In both cases, the initial
during reaction conditions. This momentum in reac- state is the zero vibrational level of the ground electronic
tion and operando Raman spectroscopy advances a new state and the final state is the first vibrational level
revolution in the catalytic Raman science that is poised of the ground electronic state. IR absorption achieves
to afford accurate molecular understanding of catalytic this state change in one step (one-photon process),
structure–activity/selectivity relationship. whereas Raman scattering requires two steps involving
photon energies the difference of which is the energy
of the vibrational transition (two-photon event). The
2 RAMAN SPECTROSCOPY photon excites the molecule, which may relax in different
ways, returning to its original energy state or to a
different energy state. The released energy equals the
2.1 Raman Effect
absorbed energy if the molecule returns to its original
The Raman effect results from the inelastic scattering of energy state (elastic or Rayleigh dispersion). One out
electromagnetic radiation upon interaction with matter, of a million times, the molecule releases an amount of
the energy exchange is in the vibrational region. Raman energy different to that absorbed (inelastic or Raman

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
4 RAMAN SPECTROSCOPY

n=2 Excitation line


(Rayleigh, at n0)
n=1 E1
1 000 000

Relative intensity (a.u.)


n=0
s
ke kes
eig
h Sto ti-sto Stokes
Rayl An 1
n0−n Anti-Stokes

n0+n

0
n=2 Frequency (cm−1)

n=1 E0 Figure 3 Schematic plot of a Raman spectrum.

n=0
Infrared absoroption Raman (trace amounts). It can be so intense that fluorescence
(one-photon event) (two-photon event) usually overwhelms the Raman signal.
Figure 2 Energy level illustrating transition in IR and Raman.
2.2 Raman Intensity versus Excitation Wavelength
dispersion). Typically, the molecule ends in a more The Raman phenomenon produces the same Raman
excited energy level. Thus, released energy is lower bands with respect to the excitation wavelength.
than the absorbed energy, which is reflected in a larger However, the use of different excitation radiations has
wavelength (lower wavenumber). This kind of Raman an effect, mainly on the Raman signal intensity and
dispersion is named Stokes. Conversely, if the final energy on the fluorescence interference. The Raman signal is
level of the molecule is lower than the original one, the directly proportional to the fourth power of the excita-
molecule releases more energy than it absorbs. Thus, the tion frequency. Shifting from IR to UV excitation line
wavelength decreases and the frequency increases. This triggers Raman scattering by almost 3 orders of magni-
dispersion is named anti-Stokes. tude. However, the Raman signal does not run parallel to
The vibrational states of a molecule depend on its such an increase. Different excitation lines have different
temperature, therefore, the chances for a molecule to features. The absorption of the radiation is different
be in a more relaxed or more excited state depends at different wavelengths; the strong absorption by the
on temperature. Therefore, the possibilities that a sample in the UV reduces the number of scatterers under
molecule be in a more relaxed or more excited state UV excitation lines.(77)
is determined by the Maxwell–Boltzmann distribution of Raman scattering may increase by several orders of
states (Equation 1): magnitude due to resonance Raman (RR) and surface-
enhanced Raman spectroscopy (SERS) effects. Very
Excited − population intense Raman bands due to resonance effect result when
= e−E/kT (1)
Relaxed − population the excitation line possesses an energy that coincides
with that of electronic state transition. Raman scattering
Stokes transitions are more likely than anti-Stokes is sensitive to all the excited electronic states of the
transitions at moderate temperatures. A Raman spectrum molecule. If the incident photon energy approaches the
shows the intensity versus the frequency of the radiation, transition energy of an excited electronic state, the Raman
as illustrated in Figure 3. The positions of Stokes and scattering changes from normal Raman scattering to RR
anti-Stokes bands with respect to the Rayleigh band are scattering. There are two important scenarios. One is
identical since it corresponds to the same transition, in the far-from resonance (FFR) limit. This is a common
different directions. Temperature affects the intensity situation for colorless samples that have no electronic
ratio between Stokes and anti-Stokes bands; anti-Stokes states in close proximity to the incident photon energy,
bands are significantly less intense than Stokes bands at which is the case in most of the spectra. The other is the
moderate temperatures. Almost all of the Raman spectra single-electronic-state (SES) limit of strong resonance
described in the literature plot Raman shifts toward lower with an SES. In this limit, the incident photon energy
frequencies from the exciting line (the Stokes–Raman is very close to, or falls within, the absorption band
spectrum). The sample may also relax by emitting of an excited electronic state of the molecule and the
fluorescence, which is about 4 orders of magnitude resulting RR scattering is dominated by the properties
more intense than the Raman signal. Fluorescence may of this resonant electronic state.(78) The Raman signal
originate from the sample itself or from some impurities also increases due to electronic surface enhancement

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 5

phenomena. The nature of SERS is not fully understood of water also absorbs the Raman signal generated upon
yet. SERS may be afforded at specific substrates, like exposure with IR radiation. In addition, the use of in situ
silver, copper, gold, photochemically roughened surface, cells for FT-Raman exhibits the same limitations as for
or graphitic materials. cells for IR spectroscopy since they must be transparent
to IR radiation. The detectors for FT-Raman, Ge, or
2.3 Benefits and Limitations of the Different InGa–As, refrigerated by liquid nitrogen afford lower
Excitation Lines in Raman Spectroscopy signal-to-noise ratio than CCD used in visible-Raman
equipments.
UV Raman affords very high Raman signal and quite often UV Raman is particularly interesting in catalysis, but
is not affected by fluorescence limitations since it works it has to be handled with care. It is attractive because
in the UV spectral window, away from the fluorescence of its ability to circumvent fluorescence problems(79)
spectral window. There is no limit on the temperature and also because of the possibility to afford deeper
of the sample. Blackbody radiation would not emit UV structural characterization through selective resonance
radiation until extremely high temperature, far above enhancements.(80) Resonance effects must be considered
the most demanding applications. However, UV Raman carefully since they change the relative intensities
possesses lower spectral resolution and, good spatial of Raman bands. The groups of Stair and Wachs
resolution. UV-Raman systems require quartz optics. have evaluated resonance effects in UV- and visible-
In addition, CCD (charge-coupled device) detectors are Raman spectra of supported oxides.(77) This comparative
not sensitive to UV-photons. Thus, CCD detectors in study comprises the evaluation of five supported metal
UV-Raman system need a special coating that converts oxide systems (alumina-supported chromia, vanadia, and
UV-photons into visible photons, detectable by CCD molybdena as well as titania-supported molybdena, and
array detectors. It is particularly important to protect the rhenia) using 514-nm (visible) and 244-nm (ultraviolet)
sample from degradation due to the large energy of the excitation lines. The UV- and visible-Raman spectra of
radiation; furthermore, interpretation of the spectra can hydrated and dehydrated alumina-supported molybdena
be difficult due to resonance enhancement effects. at monolayer coverage is presented in Figure 4. The
Visible Raman is the most common configuration. hydrated molybdate species exhibit the Mo=O Raman
Optics in this system have no special requirements and it is mode at 945 cm−1 in UV-Raman spectrum and at
suitable to study samples that exhibit weak Raman signals; 949 cm−1 in the visible-Raman spectrum. The Raman
besides, the degradation of the sample is less important band of the Mo–O–Mo bending modes appears at
than under UV excitation lines. However, fluorescence 335 cm−1 in the UV-Raman spectrum and at 210 and
problems may sometimes be problematic. There is not a 352 cm−1 in the visible-Raman spectrum. This sample
very important limitation with respect to the temperature absorbs at 244 nm in the UV–vis region.(77) The
of the sample, since it must be close to 800–900 ° C to have absorption band at ca. 235 nm is due to charge transfer
problems with strong emission radiation, which becomes from oxygen to the metal cation. The 560-cm−1 band is
visible at this temperature and overwhelms the Raman stronger in the UV-Raman spectrum and the 865 cm−1
signal. is more prominent in the visible-Raman spectrum.
NIR (near infrared)-Raman, typically under 785-nm These spectral variations appear owing to the resonance
laser is an intermediate solution between visible Raman enhancement of different bands under UV excitation.
and FT (Fourier transform)-Raman, which is very The UV- and visible-Raman spectra of the dehydrated
convenient for many organic-phase-related analyses sample are similar. The band at 1000–1003 cm−1 assigned
(pharma, biomedical), where the Raman signal remains to terminal Mo=O bond shows little dependence on
high enough and fluorescence is minimized. At NIR the excitation lines. The Raman bands from Mo–O–Mo
excitation lines, the system still takes advantage of the bending vibrations of surface polymeric species are below
convenient optical properties and glass, quartz, and fiber 400 cm−1 . The Raman bands of Mo–O–Mo symmetric
optics fit these systems, unlike FT-Raman systems, which and asymmetric stretching are in the 500–900 cm−1
use IR radiation. NIR excitation lines are becoming interval, while the O–M–O symmetric stretch is in the
increasingly popular. range 935–940 cm−1 . The difference in relative intensities
FT-Raman exhibits the lowest Raman signal. They of the Raman bands due to the symmetric and asymmetric
are very fast since they may use FT due to the fact stretches of Mo–O–Mo groups on the two spectra may
that they work with IR radiation. They afford the best be attributed to different extents of bending vibrations
spectral resolution and the worst spatial resolution. The of Mo–O–Mo. A key effect for most samples is that the
sample temperature cannot be above 150 ° C due to the strong absorption by the sample in the ultraviolet region
strong emission radiation background (IR wavelength significantly reduces the number of scatterers contributing
radiation). Water absorbs IR radiation; thus, the presence to the Raman spectrum; this strongly attenuates the

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
6 RAMAN SPECTROSCOPY

1000 1003

935

940
Dehydrated Dehydrated 865

570 206 307


Intensity

320 370

Intensity
945

952

Hydrated
Hydrated
216 865
335 560 355
560

200 400 600 800 1000 1200 200 400 600 800 1000 1200
(a) Raman shift (cm−1) (b) Raman shift (cm−1)
Figure 4 Comparison of UV Raman (a) and visible Raman (b) of hydrated and dehydrated 18% MoO3 /Al2 O3 (monolayer).
(Reproduced from Ref. 77.  American Chemical Society, 2001.)

expected larger Raman scattering cross section due • Typical spatial resolution for Raman microscopy
to resonance enhancement.(77) The UV spectra appear is ∼1-µm region, which is wavelength dependent,
to be more sensitive to the out-of-plane bending and and the spatial resolution can be further reduced
symmetric stretching vibration of bridging oxygen species to ∼10–40 nm when combined with atomic force
(M–O–M), whereas the visible-Raman spectra are more microscopy (AFM) or tip-enhanced Raman spec-
sensitive to terminal oxygen vibrations (M=O).(77) This troscopy (TERS).
underlines the advantageous complementary information • For supported catalysts, the oxide supports give
obtained by the use of different excitation lines. rise to relatively weak Raman vibrations and allow
acquisitions in the critical 700–1100 cm−1 region. In
2.4 Raman Benefits versus Infrared the case of Al2 O3 and SiO2 supports, the Raman
Compared to IR, Raman exhibits several advantages: spectra can even be collected down to ∼100 cm−1 free
of significant Raman bands interference.
• Raman may use any excitation wavelength (UV to • The gas phase usually does not contribute to
IR). vibrational spectra since light scattering from solid
• The optics in a Raman system are rather simple, in materials dominates the scattering process.
particular, for visible Raman.
• The design of in situ cells is much more versatile
than for IR if Raman works with UV or visible light.
Characterization of supported metal oxides by Raman
Quartz and glass are appropriate and the system has no
and IR demonstrates the critical difference imposed by
limitation regarding the temperature of the sample,
as long as it is not high enough to produce visible the different selection rules for Raman and IR. The
radiation (blackbody radiation, near 900 ° C). supported metal oxides consist of active metal oxide
• Microscopy accessories allow Raman studies with phases dispersed on the oxide supports, which represent
extremely little amount of sample, next to picograms the major catalyst component. Oxide supports exhibit
and low laser power on the sample. intense IR absorption bands while they are weak in
• The presence of water is not a problem if Raman Raman (silica or alumina) or show bands below 700 cm−1
works with visible or UV radiation. Thus, it is (titania, zirconia, ceria, etc). The 1100–700 cm−1 is an
possible to study aqueous solutions and system under important window for the characterization of supported
supercritical conditions. oxides (e.g. molybdena, chromia, vanadia, rhenia, tungsta,
• Raman spectroscopy is very sensitive to microcrystals etc.). Since IR spectra of supports such as alumina and
and amorphous phases, <4 nm, which are not silica exhibit extremely strong bands below ca. 1100 cm−1 ,
detectable by XRD. the important diagnostic vibrations arising from the active
• The phase of the sample is not critical (gas, metal centers, usually under 1000 cm−1 , are generally
liquid, solid). obscured.

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 7

2.5 Raman Limitations versus Infrared Other excitation lines: Using different excitation lines,
IR excitation (1064 nm), which also decreases Raman
Raman spectra have typically been more expensive than signal intensity, can be effected. This option is not
IR systems, but the price difference is decreasing. General convenient for the study of compounds with very weak
limitations are as follows: Raman signals; the use of NIR excitation (typically,
785 nm) is a rather convenient solution that minimizes
• There is a lower spectral resolution for the Raman fluorescence and does not dramatically decrease the
system, but recent Raman systems provide high- Raman signal; the use of UV excitation (below 400 nm)
resolution as an option. increases the Raman effect, but the Raman signal is lower
• Raman spectra are more difficult to quantify since owing to a much shallower sampling volume(77) and the
their intensity depends on radiation dispersion, much lower sensitivity of detectors to UV-photons. For
which depends on many parameters. It is, therefore, materials that exhibit resonance phenomena, however,
necessary to incorporate an internal reference for the spectra may be difficult to evaluate. With UV-
quantitative analyses. Raman instrumentation becoming common in the market,
• The laser may heat the sample during Raman experi- however, degradation of the sample and understanding of
ments, which may affect the structure of the analyte, resonance enhancement effects require special attention.
its temperature, induce desorption/decomposition of
probe molecules, etc. However, the use of rotors, Fluorescence depends on the energy of the incident
programming of the beam movement or fluidized beds radiation; therefore, it will be lowest under IR excitation
to distribute the beam incidence may minimize this and maximum under UV excitation. A qualitative
problem. This is an especially serious problem with representation is given in Figure 5. Excitation with UV
UV-Raman experiments that employ very energetic radiation results in very intense fluorescence; however,
UV excitation. this has little effect on the quality of the UV-Raman
• Fluorescence is a serious limitation for Raman spectrum since Raman bands are located around UV
spectroscopy, since its intensity is ∼10 000-fold more excitation radiation, while fluorescence emits in the
intense than the Raman signal. Thus, tiny amounts visible region of the spectrum, thus they do not overlap
of fluorescent impurities may overwhelm the Raman (Figure 5). Thus, the use of UV excitation lines minimizes
signal of the analyte. In some cases, the fluorescence fluorescence problems, even when it produces much more
problem may be circumvented by any of the following fluorescence than visible excitation.
steps:

Calcination: Calcination at 400–500 ° C is suitable if


the samples are stable in air at this temperature. This Fluorescence
method removes fluorescence originating from ambient
carbonaceous deposits that adsorb on the surface of
samples. This is particularly important for materials with
specific surface areas of ca. 150 m2 g−1 and above. This
treatment burns the carbonaceous species.
an
am
R
V-

Photobleaching: This would saturate the levels and


U

would minimize fluorescence emission. Photobleaching


an

is achieved by irradiating the sample with the laser prior


am

to Raman acquisition.
R
e-
bl
si
Vi

Millisecond shutter: Since fluorescence is slower than


Raman effect, an option is to run spectra at high
an
am

shutter rates, so that irradiation and Raman acquisition


-R
IR

is achieved within milliseconds, before fluorescence


N

emission occurs. Owing to short acquisition times, this


would only be valid for samples with strong Raman
signals. UV region Visible region NIR region
Absolute wave numbers
Pulsed laser excitation: This can be carried out with gated Figure 5 Qualitative plot illustrating excitation line effect on
detectors. Raman spectra versus fluorescence.

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
8 RAMAN SPECTROSCOPY

2.6 ‘‘Mass Effect’’, Isotopic Substitutions independent oscillator, vibrationally independent from
the rest of the molecule or the crystal lattice. The diatomic
The vibrational modes of atoms in a molecule or in a unit
approximation does not provide a vibrational mode
cell of a crystal can be described as harmonic oscillators
analysis, since it neglects the interaction with the nearest
associated with appropriate effective force constant and
neighbor; however, it is very efficient in determining
effective masses, which are the reduced masses (µ) –
metal–oxygen bond lengths from Raman stretching mode
i.e. for masses m1 and m2, the reduced mass, µ, is
wavenumbers.(82 – 87) The expressions that relate M–O
(m1 · m2)/(m1 + m2), of the atoms involved. In the case of
bond length to Raman stretching mode wavenumber are
a simple diatomic molecule, the vibrational wavenumber
as follows (Equation 4):
is given by Equation 2:

ν = A × exp(BR) (4)
1 K ν 1 E
ν̃ = = = = (2)
2π c µ c λ hc
in which A and B are fitting parameters, v is the
in which K is the force constant, µ is the reduced mass, Raman stretching mode wavenumber (in cm−1 ) and
c is the velocity of light, ν is the frequency, λ is the R is the metal–oxygen bond length (in angstrom).
wavelength, h is Plank’s constant, and E is the energy The A and B values have been calculated for V,(84,85)
difference between two states. An increase in mass leads Mo,(82,83) Nb,(85) W,(87) and P(88) oxides.(89) Figure 6(a)
to a decrease in wavenumber (Equation 3): illustrates this correlation for molybdenum oxide refer-
ence compounds.(83) The diatomic approximation may
 also determine the relationship between bond order
ν̃ µ
= (3) and the M–O stretching frequency(90) ; this is illus-
ν µ
trated in Figure 6(b). This correlation reveals that
In the case of the substitution of one atomic species by Mo–oxygen vibrations near 1000 cm−1 correspond to
another of a different chemical nature, the ratio of the terminal Mo=O bond, those near 600 cm−1 to single
new to the old wavenumber will, thus, be approximately Mo–O bond and those below 400 cm−1 to partial or
equal to the square root of the inverse ratio of the reduced long bonds. This correlation is most accurate above
masses.(81) If the force constants vary considerably, the 800 cm−1 , where stretching frequencies dominate.(58) A
situation becomes complex. shift of ∼10 cm−1 reflects changes of ca. 0.01 Å in the
terminal M=O bond length. An attractive aspect of the
diatomic approximation is that it can also determine
2.7 The Diatomic Approximation
the coordination and the metal–oxygen bond length
The diatomic approximation is a general approach for in materials that are diffraction amorphous. This is of
estimating the metal–oxygen bond lengths and bond particular interest in the study of dispersed oxides, like
orders from the Raman spectra of transition metal oxides. monolayers of vanadia, molybdena, etc. on different
Each metal–oxygen chemical bond is assumed to be an supports, or to the study of bulk amorphous materials

1100 1100

1000 1000
Correlation plot Mo – O
900 Mo – O bonds 900 valency Mo = O
800
800
700
700
cm−1
cm−1

600
600
500 Mo – O
500
400
400 300
300 200 Mo--O
200 100
1.6 1.8 2.0 2.2 2.4 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2
Interatomic distance (Å) Bond order

Figure 6 Mo–O bond length versus Raman stretching frequency relationship (a) and Mo–O bond stretching versus Raman
stretching frequency relationship (b). (Reproduced from Ref. 83.  John Wiley & Sons, Ltd, 1990.)

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 9

such as surface molybdenum oxide or niobium oxide scattering rejection; they are used in most Raman
species on alumina(83,85) or amorphous bulk oxides.(87) microscopes. These filters typically reduce the intensity
of a small wavelength range containing the laser line by 8
orders of magnitude while transmitting 100% of all other
3 RAMAN SYSTEM wavelengths. Typically, these filters allow Raman spectra
at Raman shifts above 80 cm−1 .
Raman spectrometers are constituted by an excitation
source (laser), sample illumination, and collection of 3.2 Detectors
scattered radiation. The collected radiation is taken to Because of the weak Raman signal, the main problem is
the spectrometer, but elastic scattering must first be the detection of a significant amount of Raman photons.
rejected. Originally, the detection was photographic; it has been
replaced by photomultipliers, multichannel detectors,
3.1 Excitation and Laser-line Rejection in CCD detectors, and those specific for FT-Raman(91,92)
Spectrometer CCD detectors are the most popular detectors nowadays.
They consist of bidimensional arrangements of pixels
Raman lasers typically work in the UV, visible, NIR,
with each pixel ∼6–30 µm. They possess high quantum
and IR. It is necessary to remove the elastic disper-
efficiency and sensitivity in a wide spectral window
sion radiation (Rayleigh) from the inelastic one (Stokes)
(120–1000 nm). The very high sensitivity of CCD
since the former is more than 6 orders of magnitude
detectors makes them sensitive to interference caused by
stronger that the latter. The detectors are then ready
cosmic rays. Cosmic rays are atomic nuclei, stripped from
to register the Raman signal. There are several options
electrons. When they pass through a CCD detector, they
to remove the elastic dispersion, which determine the
generate sharp peaks in the Raman spectra. Most Raman
resolution and the signal-to-noise ratio. Traditionally,
system have a software option to discriminate cosmic
single-monochromator configurations disperse the radi-
rays, typically based on comparing several spectra.
ation but did not provide an efficient system due to
extraneous light limitations. The use of two monochro-
mators afforded an efficient line rejection. The use of 3.3 Calibration in Intensity and Wavelength
triple-monochromator configuration is limited to very Raman systems require calibration of the frequency
specific applications demanding high spectral resolution and intensity readings.(91,92,93) Raman shift calibration
and the need to study Raman shifts very close to the is particularly important when resolution must be close to
elastic scattering (below 50 cm−1 ). 1 cm−1 and an internal reference must be included with
The configurations described above are expensive the sample, like solvent Raman bands, or an admixed
and decrease Raman signal. Thus, conventional triple- solid. A calibration of the system prior to operation will
monochromator Raman systems require up to 5 W. be sufficient if resolution need not be better than 1 cm−1 .
Rayleigh filters make it possible to use a single Typical references are laser plasma lines or compounds
monochromator (Figure 7). Notch filters have good with intense sharp Raman bands, like silicon wafers or
transmission of Stokes and anti-Stokes Raman bands, diamond.
while edge filters have good transmission on the Raman is typically not quantitative; intensity calibration
Stokes–Raman bands side only. These filters are strongly depends on the laser power on the sample, on the
becoming increasingly efficient and sharp for elastic wavelength of the excitation line, on the spectrometer,
on the collection of the scattered radiation, on the
radiation absorption of the analyte, on its concentration
El jec
as tio
re

Dispersion and aggregation state, on its texture, on its color, etc.


tic n
di filte

grating D tor
sp r

CC tec Thus, quantitative Raman analyses do require an internal


er
si

de
on

reference to normalize Raman intensity values.


or
irr
M

Slit Prismatic mirror 3.4 Optical Sampling System


Illumination/
Spectrometer The two most common configurations for Raman
collection
lens acquisition are called ‘‘90° ’’ and ‘‘180° ’’, which indicates
Laser
the angle between the incident beam and the direction
to the radiation collection. Classical sampling at 90°
M
irr
or

Sample (macrosampling) requires significant efforts to align the


Figure 7 Scheme of a single-monochromator Raman system. position and orientation of the sample to direct scattered

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
10 RAMAN SPECTROSCOPY

radiation essentially to the spectrometer. A 180° sampling which is Latin for ‘‘working’’ or ‘‘operating’’, is
is simpler since it depends only on the distance. used to emphasize the simultaneous evaluation of
Fiber-optic Raman probes direct the scattered Raman both structure and catalytic performance,(3,23,66,68,69)
light to the spectrometer. There is no need for alignment including use of a cell that delivers reaction
and additional optics. Fiber-optic probes are increasingly kinetics data that match those obtained in an ideal
being used for in situ Raman studies owing to their ease of reactor. Because operando spectroscopy requires
use and convenience. Raman probes may work at distant the combination of a spectroscopic technique and
positions from the spectrometer. This is convenient for activity measurement equipment, this should reflect
dangerous or extreme processes. The work by Slater in the nomenclature; e.g. Raman spectra during
et al.(94) carries a complete discussion about fiber-optic catalytic operation with simultaneous online gas-
Raman probes. Some catalytic applications have been chromatograph activity measurement should be
reported elsewhere.(22,44) named ‘‘operando Raman-GC’’.

The in situ Raman measurements described above are


particularly valuable for catalysis since they are time
4 RAMAN SAMPLING FOR
resolved, which provides a direct feedback on molecular
CHARACTERIZATION OF SAMPLES IN transformation while the reaction proceeds.
REACTIVE ENVIRONMENTS

Raman spectroscopy applied to catalysts in controlled and


reactive environments are often described as ‘‘in situ’’. 5 EQUIPMENT FOR RAMAN
The term ‘‘in situ’’ is Latin for ‘‘on site’’, i.e. the sample SPECTROSCOPY OF CATALYSTS IN
is analyzed at the location where it has been treated or REACTIVE ATMOSPHERES
is being treated; etymologically, in situ has no temporal
discrimination. Several levels of such experiments are Since the early Raman times, several groups have made
described here: cells to control pressure, temperature, and reactants’
environment. Grasselli et al.,(34,101) described a heated
1. ‘‘In situ’’ spectroscopy: The spectra of a sample are cell used to follow transformations in bulk molybdate
recorded at the same location at which it has been, catalysts. Designing cells has became much simpler
or is being, treated; the temperature or gas phase, with the Raman microscopes, which allow working
however, may be different. in a backscattering mode (180° ).(76,102 – 107) A cell with
2. Variable-conditions ‘‘in situ’’ spectroscopy: Trans- a Raman microprobe was used to monitor phase
formations occurring during the variation of a transformations of CoMoO4 and NiMoO4 upon heating
parameter, such as partial pressure of a component, in air and to follow the sulfidation of g-alumina-supported
temperature, etc. are monitored spectroscopically. NiMoO4 catalysts at 320 ° C in mixtures of H2 and
Typical are temperature-programmed processes(95,96) H2 S.(50,108) Commercially available Raman microscope
such as TPR (temperature-programmed reduction)- cells often are not designed for the characterization of
Raman spectroscopy, in which Raman spectra catalysts in reactive environments, and problems may
characterize the reduction of a sample,(97,98) TPO arise from temperature inhomogeneity and limited gas
(temperature-programmed oxidation)-Raman diffusion. The investigation of liquids, gases, or solids in
spectroscopy,(97,99) or any temperature-programmed reactive atmospheres requires the use of vials or cells that
reaction with an adsorbate or a probe molecule.(100) have to be transparent to the radiation. Glass or quartz
3. ‘‘In situ’’ spectra under conditions relevant to catalytic is suitable for the construction of cells, making Raman
operation: The spectra are taken under conditions spectroscopy very versatile, particularly when compared
relevant to catalytic operation. Online analyses (mass to IR spectroscopy.
spectroscopy (MS), gas chromatography (GC)) assess
if the spectra correspond to a catalyst at work.
5.1 Reactor Cells for Raman Spectroscopy in the
However, the design of many in situ cells is not
Liquid Phase
appropriate to afford true catalytic data.
4. Operando spectroscopy: This involves ‘‘in situ’’ Increasing numbers of Raman investigations aim at
spectroscopy of the working catalyst. To demonstrate solution chemistry processes. One example of an experi-
that the spectra correspond to an operating catalyst, mental design for solution Raman spectroscopy is
quantitative analysis of the reaction progress (e.g. by shown in Figure 8. The respective precursor solutions
GC) has to be performed, and in this case, structure containing various concentrations of transition metal ions
and activity can be correlated. The term ‘‘operando’’, are pumped through a cuvette with a plane parallel

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 11

Spectrometer
and selectivity values were not measured simultaneously
Detector with the spectra. The developments of these Raman
experiments have been reviewed elsewhere.(22,24,37,122)
One of the first cells investigated the structural evolu-
Laser tion of alumina-supported molybdenum oxide catalysts
during calcination was developed by Schrader et al.(118)
(Figure 9a); the key feature of the design is the place-
Pump ment of the catalyst, which is pressed into a wafer on
Cuvette a rotor. Chan and Bell(117) reported on the formation
of silica- and lanthana-supported palladium catalysts in
this kind of reaction chamber. Cells using a rotating
sample to prevent sample degradation in the beam
have been described in detail.(35,36,114 – 118,123) Figure 9
illustrates the concept of Raman cells for reactive envi-
pH ronments. Several commercial cells are suitable for use
Temperature control
control in combination with Raman microscopy. The rotating
Figure 8 Concept scheme of an experimental design for sample design was modified by Wachs’ group (Figure 9a)
solution Raman spectroscopy. and used to investigate supported oxides during selective
alkane oxidation(124,125) and methanol oxidation.(100,126)
Figure 9(b) shows a Raman cell similar to that developed
optical window. Both the incident beam and the Raman
and reported by Lunsford et al.(127) The void volume
scattered light pass through this window. The pH value
upstream of the powder catalyst is minimized, and the
is monitored by a pH meter in the storage vessel, and a
gases flow through the bed. The cell was used for investi-
controller is used to adjust the solution temperature.
gating catalysts for NOx removal,(120,121,128 – 130) including
Titroprocessors are used to control the addition of
the determination of kinetics. To make possible the corre-
further precursor solutions or of the pH by automated
acid or base addition.(109) In 1982, Woo and Hill(110,111) lation of spectroscopic and catalyst performance data, the
reported an investigation of Co2 (CO)8 and (RuCl(CO)2 )2 temperature of the analyzed spot has to be representa-
during propylene hydroformylation, with the complexes tive of the catalyst bed. Thus, local heating by the laser
in solution or supported on silica or alumina. They used a beam has to be minimized. Volta et al.(113) designed a cell
Raman cell autoclave reactor, which was combined with similar to that in Figure 9(b), including a lens rotating
online GC analysis of the effluent stream. This successful off axis, thus avoiding local heating at particular spots.
attempt to use a spectroscopic cell as a reactor represents Other lens systems distributing the laser light have also
the earliest Raman investigation of a working catalyst been described.(131 – 133) Stair’s group(79) has developed
with simultaneous product analysis.(112) a cell for UV–Raman spectroscopy similar to that in
Figure 9(b) but where the fixed bed is replaced by a
fluidized bed. Futhermore, to facilitate tumbling of the
5.2 Cells for Raman Spectroscopy of Solids in Reactive catalyst particles and thus to maintain the fluidization, the
Environment Raman cell was mounted on a shaker.
Raman spectra of catalysts in a controlled environment Experiments in which catalyst wafers are used may
may be acquired over a wide range of temperatures and suffer from reactant mass transfer problems, which limit
pressures. The upper temperature limit depends on the the validity of the data or complicate their analysis. To
excitation wavelength, which determines the range of determine the reaction kinetics and activation energies,
sensitivity of the detector. Blackbody radiation in the mass transfer effects have to be understood.(134) These
same range can contribute significantly to the detector difficulties can be avoided if a conventional fixed-bed
signal and can obscure spectral information. Activity reactor is mimicked closely and the catalyst is used
and kinetics data and activation energy values measured in powder form and the reactant gases flow through
with such a spectroscopic cell have to be consistent the bed. It is also important to prevent homogeneous
with those obtained with corresponding conventional gas-phase reactions by reducing the dead volume. A
catalytic reactors. Raman spectroscopy has been used criterion for the suitability of a spectroscopy cell for
frequently to investigate the chemisorption of probe investigations of working catalysts can be formulated
molecules.(27,64,65) Several groups reported Raman cells in as follows: the activity or selectivity data and activa-
which the temperature of the sample and the environment tion energy values have to be in agreement with the
can be controlled so that catalytic reaction conditions can catalytic performance data measured with a conven-
be simulated.(113 – 122) In these investigations, conversion tional fixed-bed reactor. Table 2 is a comparison of

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
12 RAMAN SPECTROSCOPY

To Laser beam
sp
ec
tro
m Quartz window
et
er
Oven

Quar Catalyst
t
wind z Inlet
fixed bed
ow
Rotor Oven
Thermocouple Outlet
Outlet

am
r be Inlet
se
La
Catalyst wafer Thermocouple
(a) (b)
Figure 9 Illustration of the concepts of Raman cells that can be used for experiments under reaction conditions: (a) similar to the
one developed by Schrader et al.(118) ; (b) similar to the one developed by Lunsford et al.(127)

Table 2 Catalytic activity of a catalyst of molybdenum and vanadium oxides on alumina with a Mo : V
atomic ratio of 1 : 1 at monolayer coverage during propane ODH,(135) measured in a conventional
fixed-bed reactor and a fixed-bed Operando reactor cell
Reactor type Temperature ( ° C) Conversion (%) Selectivity (%)
Propane Propene CO CO2
Conventional 320 5.7 49.5 31.2 19.3
Operando 320 5.6 45.6 35.6 18.8

the conversion and selectivity values characterizing an to prepare industrial catalysts.(138,139) The Raman cell
alumina-supported molybdenum–vanadium oxide cata- designed by Mestl et al.(43,140,141) includes a fiber-optic
lyst during propane oxidative dehydrogenation (ODH) Raman probe and a double-bed configuration. This is
obtained with a conventional fixed-bed reactor and with the only reported configuration with a tubular fixed-
a spectroscopic cell that fulfills this requirement,(135) in bed flow reactor that allows the investigation of the
addition, activation energy values in the fixed-bed reactor working catalyst at steady state in contact with the
and in the operando-bed reactors are consistent.(136) reactant and product mixture as a function of gas-
Similar considerations have also been reported earlier
phase residence time.This configuration has been used
for other methods, such as X-ray diffraction.(67) The
to gain understanding of the function of bulk mixed metal
reactor cell used in an early Raman investigation by
oxides in selective partial oxidation reactions.(43,140,141)
Hill et al.(137) of bulk and supported iron molybdate
A representative Raman cell designed to meet above-
catalysts during methanol oxidation to formaldehyde
stated criteria is presented in Figure 10. The home-made
was designed to minimize void volume and temperature
gradients so that activity data would be representative operando reactor designed by Bañares et al.(142) uses a
of the catalytic process.(137) No significant structural Raman microscope system in combination with a fixed-
transformations of the catalyst were observed during bed and online GC and MS for analysis (Figure 10c).
reaction at 275 ° C. This result is consistent with those of The microreactor walls have optical quality. Remark-
other Raman investigations.(131 – 133) Experiments under able about this design is that no appreciable differences
reducing conditions (with methanol) and oxidizing condi- in activity and selectivity can be observed between
tions showed that segregated molybdena reoxidizes more the Raman reaction cell and a conventional fixed-
readily than Fe2 (MoO4 )3 .(137) More recent investigations bed microreactor.(135) The furnace is designed with a
emphasize how to take advantage of methanol-vapor- small opening that allows spectra acquisition with a
induced changes of the dispersion of the active phase long-distance objective and does not lead to a local

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 13

Raman spectra were recorded under ambient conditions,


the spectra of the catalysts with low molybdenum
loadings indicated the presence of monomolybdates, as
Window characterized by the prominent band at about 900 cm−1 .
In contrast, polymolybdates, indicated by Raman bands
Heat screen
Oven
between 930 and 960 cm−1 , were detected at intermediate
Gas inlet loadings, and crystalline MoO3 nanoparticles were found
Heated lines Thermocouple in materials with high loadings. This picture of the surface
species, however, changed when the Raman spectra
were recorded at high temperatures under controlled
Inert packing (SiC) conditions(148) ; the noncrystalline surface-MoOx species
were then characterized by a Raman band at about
1000 cm−1 . The exact nature of the surface molybdenum
Heating species under conditions of low and high humidity was
Metal cartridges
block
still under debate at time. This dynamic behavior of the
catalyst, depending on the degree of hydration, is typical
Figure 10 Fixed-bed operando Raman reactor designed by of many supported transition metal oxides, as evidenced
Bañares. (Source: M. A. Bañares).
by the changes in their Raman modes,(114,115) as discussed
below.
temperature decrease. This cell has been used to analyze
the structural transformation of supported oxide cata- 6.2 Hydrated Supported Metal Oxides
lysts during ethane oxidation,(142) propane oxidation,(66,68)
The first attempt to measure the Raman spectrum of
and propane ammoxidation.(66,69,70) Weaver et al.(143,144) supported MoO3 /Al2 O3 was reported by Trifiro et al. in
investigated NO oxidation with CO in a SERS experiment 1972.(149) The absence of detectable Raman bands led
with MS analysis and reported simultaneously recorded these investigators to conclude that the dispersed molyb-
Raman and activity data. The details of their cell design date species must be Raman inactive. In 1977, however,
were not disclosed. Brown(114 – 116) successfully reported Raman characteriza-
tion of supported MoO3 /Al2 O3 and MoO3 /SiO2 –Al2 O3
catalysts, as well as with NiO- and CoO-promoted
6 RAMAN IN CATALYSIS, A JOURNEY catalysts. Furthermore, they also observed the pres-
THAT STARTED IN THE 1970S ence of crystalline MoO3 and Al2 (MoO4 )3 in the
Raman spectra, and the latter phase was not even
6.1 Early Vacuum, Chemisorption, and detectable by X-ray diffraction (XRD). This was a break-
Hydration/Dehydration Experiments through in Raman characterization of supported metal
oxide catalysts, which was rapidly followed by publi-
Early Raman experiments characterizing catalysts(114 – 118) cations from other research groups reproducing the
took advantage of sample rotation to avoid sample Raman spectrum of supported MoOx /Al2 O3 ,(150 – 153) as
damage and desorption of molecules.(35,36,123) These early well as other supported metal oxides: V2 O5 /Al2 O3 ,(154)
investigations concerned sulfidation/reoxidation treat- Re2 O7 /Al2 O3 ,(155) CrO3 /Al2 O3 ,(156) and NiO/Al2 O3 .(108)
ments of molybdenum–cobalt oxide on silica–alumina Over the past two decades, the ambient Raman spectra
and showed the conversion of surface molybdenum oxides of numerous supported metal oxide catalysts have been
into sulfides. Partial reoxidation to oxysulfide could be reported.(22,23)
avoided by applying vacuum; this allows determining the Under ambient conditions, moisture adsorbs on the
presence of MoS2 on alumina.(114,115) This procedure also surface of supported metal oxide catalytic systems; this
allows to study fresh and used supported metal oxide cata- leads to an extensive solvation of the surface metal
lysts after various treatments.(114,115,145) A cell designed to oxides (equivalent to ca. 20–40 monolayers of water)(23) ;
rotate the powder sample and to heat it to a temperature thus, many contradictory models for the dispersed metal
of 450 ° C, in vacuum or in a selected gas atmosphere, was oxide structure were proposed. In 1983–1984, Wang and
used to investigate the chemisorption of pyridine on silica Hall,(153) Chan et al.,(147) and Stencel et al.(145) published
gel as a function of temperature(146) and also to monitor papers that supported theories that Re2 O7 , MoO3 ,
the effect of various treatments on various samples.(118) and WO3 –V2 O5 were hydrated during ambient Raman
One of the key discoveries during the first Raman measurements. However, the molecular structures of the
experiments was a demonstration of the influence of various hydrated dispersed metal oxide species on oxide
moisture on the spectra of some catalysts(147) ; when the supports were not fully understood at that time.

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
14 RAMAN SPECTROSCOPY

In subsequent years, various research teams real- in situ Raman, UV–vis–NIR DRS (diffuse reflectance
ized that the Raman spectra of the hydrated dispersed spectroscopy) and X-ray absorption near-edge structure
metal oxide species present on oxide supports resemble (XANES) spectroscopy also revealed that the structure
those of the polyoxo anions of the respective metals of hydrated surface vanadium oxide species is dependent
in aqueous solution.(150,157 – 160) Furthermore, like those in on the degree of hydration.
aqueous solution, the molecular structures of the hydrated To summarize, the surface chemistry of adsorbed
supported polyoxo anions varied with the isoelectric point metallates in the fully hydrated state of the surface
(IEP) or the point of zero charge (PZC)(161) of the oxide closely resembles their aqueous solution chemistry.
surface.(162) Deo and Wachs confirmed that the solution A wealth of Raman spectra have been published
chemistry that determined dominant Vx Oy species in the characterizing pure molybdenum, vanadium, and tung-
liquid phase, depending on concentration and pH in the sten polyoxometallates.(177 – 193) Humidity may actually
liquid phase is consistent with that observed for hydrated dissolve the support and form heteropoly anions during
supported oxides.(162) These observations allowed Deo the impregnation step(194,195) or after prolonged exposure
and Wachs(162) to predict the various hydrated molec- to moisture.(126,196 – 198)
ular structures on surfaces. At low surface coverage
of the hydrated dispersed metal oxide species, the net 6.3 Dehydrated Supported Metal Oxide Catalysts
PZC is dominated by the PZC of the oxide support.
However, the PZC is a function of the PZC of the The generalization of environmental Raman cells shifted
oxide support and the dispersed oxide component at attention to the molecular structures of dehydrated
higher surface coverage. At monolayer surface coverage, dispersed metal oxide species on supports. To the best
the PZC characterizing hydrated supported metal oxides of our knowledge, the first report of Raman spectra
appears to be approximately the mean of the values of of dehydrated species of supported MoO3 /Al2 O3 cata-
the two components. The PZC theory was confirmed for lysts calcined in O2 at 550 ° C was performed in 1980(118) ;
many supported metal oxide materials under ambient this work also reported the Raman bands of adsorbed
conditions, for which both the Raman spectra and pyridine. Subsequent Raman investigations of dehy-
the pH at the PZC were determined: V2 O5 ,(162,163) drated supported Re2 O7 ,(153) V2 O5 ,(147) MoO3 ,(145,147)
Re2 O7 ,(162,164) MoO3 ,(148,165,166) WO3 ,(167) CrO3 ,(168) and WO3 ,(145,147) CrO3 ,(199) Nb2 O5 ,(200) TiO2 ,(201) ZrO2 ,(202)
Nb2 O5 .(169) The Raman investigation of niobium species Al2 O3 ,(203) and Ta2 O5 ,(204) demonstrated that dehydrated
in aqueous solutions of niobium oxalate(170) nicely surface metal oxide species had unique molecular struc-
showed the dependence of their constitution on pH tures that were dependent on both the specific oxide
and concentration. The PZC theory was successfully support and the surface coverage of the dispersed
applied to predict the hydrated, molecular structures metal oxide species. These results can be explained
of multicomponent supported metal oxide species, by the condensation reactions of the surface metal-
such as iron–molybdenum, iron–vanadium, molyb- late ions with the support surface OH groups, which
denum–vanadium, tungsten–vanadium, and sodium– were confirmed by IR spectroscopy.(165,166) The complete
vanadium oxide species.(171,172) PZC theory for the resolution of the structure required the combination
prediction of the molecular structure of hydrated polyoxo of Raman spectra of the dehydrated dispersed metal
anions holds true only under ambient conditions when oxides on supports with data obtained with complemen-
the oxide surfaces are extensively hydrated. This condi- tary methods: XANES/extended X-ray absorption fine
tion is not satisfied when the supported metal oxide structure (EXAFS), solid-state NMR, IR, and diffuse
catalysts are heated to elevated temperatures (>100 ° C) reflectance UV–vis spectroscopies, and isotopic 18 O-
and physisorbed water desorbs. Even in the pres- exchange experiments. On SiO2 , dehydrated surface
ence of flowing steam, dispersed metal oxide species metal oxide species were found to be generally present
essentially become dehydrated at elevated temperatures as isolated molecular structures,(172) because of the
(230–500 ° C), because of the rapid desorption of water low density of the reactive anchoring surface Si–OH
at such temperatures.(173) Recent Raman experiments sites with which adsorbed metallate ions condense
carried out at temperatures up to 500 ° C with samples in (releasing water). Polymeric species would exhibit Raman
the presence of water confirmed this trend.(174,175) bands originating from bridging M–O–M bonds: d
For the partially hydrated sample, the broad Raman (ca. 200–300 cm−1 ), ns (ca. 450–550 cm−1 ), and nas
bands observed are very different from the bands (ca. 650–750 cm−1 ). These bands were absent from
observed for the fully hydrated and dehydrated the spectra of dehydrated silica-supported metal oxide
samples.(176) Therefore, the molecular structure of the species. On other oxide supports, dehydrated surface
surface vanadium oxide species on silica is also depen- metal oxide species were detected as both isolated and
dent on the degree of hydration. A combination of polymeric molecular structures.(23) Isolated surface metal

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 15

oxide species predominate at low surface coverage, and function of vanadia coverage suggest that the surface
polymeric surface metal oxide species tend to be the VO4 species also polymerize on these supports. Analo-
major surface species at high surface coverages. gous UV–vis studies for tungstates,(213) molybdates,(214)
The molecular structures of the dehydrated group chromates,(205) niobates,(215) and rhenates(205) have also
5–7 supported metal oxides are characterized by been reported.
the presence of oxo groups, generally appearing at Acidic surface metal oxides with cation oxidation
∼1000 cm−1 , and several bridging oxygen bonds to states of 4–7 anchor to oxide support by reacting with
the support or other neighboring cations.(23) For the the surface hydroxyl groups, but basic surface metal
supported group 5 metal oxides, only mono oxo, M=O, oxides with cation oxidation states of 1–3 anchor to
surface structures are present.(205,206) The supported oxide supports at surface Lewis acid sites.(169,216 – 219)
group 6 metal oxides, however, are present as both Raman spectra demonstrated that supported basic metal
monoxo and dioxo, O=M=O, surface structures.(206) oxides are insensitive to moisture. The Raman spectra
Fortunately, the vibrational range for both structures of basic surface metal oxide species do not show the
are quite different that allows their discrimination (e.g. bands at about 1000 cm−1 that would indicate terminal
surface O=W=O species vibrate at ∼980 cm−1 and M=O bonds. The spectra typically exhibit Raman bands
surface W=O species vibrate at ∼1015 cm−1 ). The surface in the 500–700 cm−1 region, characteristic of M–O
rhenium oxide species was the only group 7 system bonds.(212,219,220) Similar behavior was also observed for
examined and it was always found to be present as a dispersed surface TiOx, ZrOx, PtO2, and other oxide
trioxo surface structure (Re(=O)3 ).(205) The additional surface species with cations in the +4 oxidation state. Gao
vibration at ∼910 cm−1 has been assigned to bridging et al.(221) performed combined Raman and UV–vis DRS
M–O-support bonds on the basis of complementary IR analyses of TiO2 /SiO2 and found that the coordination
and Raman data characterizing supported oxides and of the surface TiOx species varied from isolated TiO4
model catalysts as well as density functional theory to polymeric TiO5 to TiO2 (anatase) nanocrystals with
(DFT) calculations.(40,68,207 – 210) When bridging M–O–M increasing titania loading. The relative intensity and
bonds are present, they give rise to vibrations at positions of the Raman bands of the TiO2 (anatase)
νas (∼650–750 cm−1 ), νs (∼450–550 cm−1 ), and bending nanocrystals did not depend on the crystallite size in
(∼200–300 cm−1 ).(23) The M=O vibrations also slightly the 2–25 nm size range.
shift to higher wavenumber with surface coverage, which
is related to the polymerization of the surface MOx
6.4 Hydrated and Dehydrated Bulk Metal Oxide
species with surface coverage (see discussion in the
Catalysts
following paragraph). The Raman bands of the dispersed
surface oxides also shift and weaken with temperature, Raman structural information characterizing crystalline
which was attributed to thermal expansion and changes bulk metal oxide phases is typically consistent with
in the population of the vibrational energy levels, XRD. When the bulk metal oxide phase does not
respectively.(211) have long-range order (i.e. is X-ray amorphous), Raman
Combination of UV–vis, DRS, and Raman spec- spectroscopy provides structural information.(124,222 – 227)
troscopy data has allowed for the quantitative deter- Water typically does not diffuse into the bulk of
mination of the monomer and polymer concentrations of the lattice of metal oxides and Raman spectra are
the surface metal oxide species.(212) The dispersed surface identical in dry and wet environments. This lack
vanadia phase in dehydrated supported V2 O5 /SiO2 cata- of sensitivity does not exist for hydroxide-containing
lysts was found to consist predominantly of isolated structures(170) and heteropolyanions stabilized by water
surface VO4 species at maximum dispersion. Isolated molecules,(126,228 – 230) which are sensitive to humidity.
surface VO4 units in supported V2 O5 /Al2 O3 and Leroy et al.(231) reported the first Raman spectra of
V2 O5 /ZrO2 catalysts, however, were present only at bulk metal oxide catalysts (Fe2 (MoO4 )3 ) in 1971. In
coverage below 20% of a monolayer (<1.5 V nm−2 ). The subsequent years, the Raman spectra of numerous mixed
fraction of polymeric surface vanadia increases more metal oxides were reported; a chronological summary
rapidly with coverage on alumina than on zirconia. At can be found in a review by Wachs.(61) Beyond providing
monolayer surface coverage, 100% of the vanadia species bulk structural information about 3-D metal oxide
on alumina (and about 80% on zirconia) are present as phases, Raman spectroscopy can also provide informa-
polymeric surface VO4 species. Although UV–vis DR tion about the terminating (and thus bidimensional)
spectra of vanadia on other oxide supports (such as TiO2 , surface layers of bulk metal oxides. For example,
CeO2 , and Nb2 O5 ) cannot be readily interpreted because surface Nb=O, V=O, and Mo=O functionalities were
of the overlap of their strong absorptions with those detected by Raman spectroscopy for bulk Nb2 O5 ,
of vanadia, equivalent shifts of the Raman bands as a and for vanadium–niobium, molybdenum –vanadium,

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
16 RAMAN SPECTROSCOPY

molybdenum –niobium, and vanadium–antimony mixed and octamolybdate, and decavanadate become apparent
oxide phases.(70,170,232) upon acidification. This is in line with the trends reported
by Deo and Wachs on the role of pH for liquid-phase
chemistry of transition metal salts.(162) At the pH value of
7 OPERANDO AND TIME-RESOLVED, 1, the bands of the pure polyoxo metallates are replaced
SPACE-RESOLVED, AND VARIABLE by those of a mixed polyoxometallate species. The next
TEMPERATURE-PROGRAMMED step in the preparation of typical multimetal mixed oxide
SPECTROSCOPY catalysts after formation of the precursor solution is the
removal of the solvent by drying.
Bulk VPO catalysts pass through various synthesis
Raman spectroscopy is increasingly being used to assess
stages; in addition, they undergo further transformations
the state of materials under reactive environments
during catalytic reaction.(113,124,225,241,242) The phase
and during material syntheses. The progress with in
transformations in mesostructured VPO phases during
situ characterization has revealed that the structure
thermal treatment in N2 were monitored by Raman
of a catalyst is typically not the same under ambient
spectroscopy(243) . Raman spectra successfully monitor
conditions as under in situ dehydrated or during catalytic
the progressive decomposition of the surfactant(243) ;
operation. In addition, a number of industrial catalytic
two new Raman bands, near 1590 and 1350 cm−1 ,
systems based on transition metals are subjected to
induction periods during operation, which results in the became evident as the treatment temperature increased.
formation of phases that may be related to the observed These bands are characteristic of graphite-like carbon
catalytic activity. Thus, in many cases, confusion arose species, which were generated from occluded surfactant
from attempts to correlate the structure that a catalyst molecules. The absence of second-order Raman spectra
possesses at conditions that are not relevant to the of the carbonaceous deposits in the wave number
catalytic reaction conditions. The examples below are range of 2700–3200 cm−1 emphasizes the lack of long-
representative of the progress in the use of Raman range order in these aggregates.(244 – 246) An important
characterization. stage of catalyst preparation is the formation of the
actual active phase. Vanadium antimonate catalysts for
propane ammoxidation are activated in the initial period
7.1 Raman Spectra during catalyst synthesis
on stream. Operando Raman-GC analyses shows that
There is an increasing number of Raman spectra being catalysts consisting of surface vanadia and antimony oxide
reported during catalyst preparation (oxides, zeolites, species combine into bulk rutile VSbO4 phase during
or metals). Such experiments deliver information about ammoxidation reaction(69) ; detailed Raman analyses
molecular structures and the formation of crystalline show that the redox cycle takes place, which is discussed
phases. This is particularly convenient since crystalline below.
phases are detected at earlier stages by Raman spec- Few Raman investigations include reports of the prepa-
troscopy than by XRD. The preparation of mixed oxides ration of catalysts containing dispersed metals, such as
(233)
or zeolites(223,224,234 – 238) have been studied by Raman silica- or lanthana-supported palladium.(117) The trans-
spectroscopy. For instance, Raman spectra show that formation of the PdCl2 precursor into supported PdO
tetrapropylammonium template is initially trapped in an was observed by Raman spectroscopy; however, the exact
amorphous silica phase.(236) Raman spectra also showed form of the palladium species could not be determined,
that Al(OH)− 4 polymerized into aluminosilicates that and it was suggested that a lanthanum–palladium mixed
formed the nuclei for crystal growth.(237) A comprehen- oxide formed.
sive review about the genesis of zeolites characterized by
Raman spectroscopy is available.(42,237)
A wealth of Raman literature has been published 8 RAMAN SPECTRA DURING
characterizing the aqueous solution chemistry of TREATMENTS
pure molybdenum, vanadium, and tungsten polyoxo-
metallates.(177 – 193,239) Raman spectroscopy is also applied
8.1 Temperature-programmed Reduction-/
for the characterization of mixed metal ion solutions
Temperature-programmed Oxidation-Raman
with compositions of industrial relevance.(109) The Raman
spectra of solutions containing molybdenum, tungsten, TPR and TPO afford valuable information on the
and vanadium ions as a function of pH have been dynamic states of catalytic materials. An interesting
reported by Mestl et al.(44,109,240) At high pH values, case is the tendency of dispersed oxides to primarily
for example, between 9 and 6, Raman bands evidence tend to aggregate upon reduction. Raman spectra
MoO4 2− and V10 O28 6− species. Raman bands of hepta- during TPR of silica-supported vanadia at a third of a

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 17

the Raman band at 982 cm−1 was reduced before the


37

Raman band at 1011 cm−1 . The 982-cm−1 band corre-


10

Reoxidized (600°C)
sponds to the νs (O=Cr=O) stretch from dioxo surface
4

4
99

28

6
14
CrO4 species and the 1011-cm−1 band corresponds to
2
70 the νs (Cr=O) stretch from mono oxo surface CrO4
or CrO5 species. The relative reactivity of the two
650 °C, 10% H2 surface CrOx species during time-resolved oxygen-18
exchange with H2 18 O also showed that the surface
Temperature-programmed

dioxo CrO4 species exchanged more rapidly than the


550 °C, 10% H2
surface mono oxo CrO4 or CrO5 species.(206) Time-
resolved oxygen-18 exchange studies also allows for
determination of the number of terminal M=O bonds
300 °C, 10% H2 present for surface MOx species.(206) A similar work
reduction

was also reported by Korhonen et al.(247) for CH3 OH


interacting with supported CrOx /ZrO2 catalysts, which
contain monolayer surface CrOx coverage) (Figure 12).
Fresh dehydrated
The initial Raman spectrum of the dehydrated cata-
lyst reveals that only surface CrOx species are present
on the ZrO2 support. In addition, a series of Raman
bands are also present in the 400–800 cm−1 region from
1100 900 700 500 300
the ZrO2 (monoclinic) support. Adsorption of CH3 OH
Figure 11 Fresh dehydrated, during TPR-Raman and reoxi- at 100 ° C significantly decreases the intensity of the
dized silica-supported vanadium oxide at monolayer coverage Cr=O vibration due to the interaction of the surface
(0.8 V atoms nm−2 ). (Source, M.A. Bañares).
CH3 O∗ intermediates with the surface CrOx species,
which are readily detectable by IR. The decrease in
monolayer show a progressive decrease of the intensity intensity is due to hydrogen bonding of the methyl hydro-
of the terminal V=O bond of the surface vanadium gens of the surface methoxy with the Cr=O bonds.
oxide species at 1037 cm−1 , which runs parallel to H2 Heating in helium from 100 to 325 ° C transforms the
consumption. The Raman spectrum of the fresh catalyst surface methoxy into HCHO and water and reduces the
is restored upon reoxidation. The process is reversible surface CrOx species. Reoxidation of the reduced cata-
upon reoxidation. However, reducing environments have lyst generates the initial dehydrated Raman spectrum
an effect on vanadia dispersion when surface vanadia reflecting the complete redox catalytic cycle for methanol
coverage approaches monolayer. Silica-supported surface oxidation to HCHO and water over the supported
vanadium oxide undergoes a structural modification upon CrOx /ZrO2 catalyst.
reduction in H2 and upon reoxidation (Figure 11) new
Raman bands of crystalline V2 O5 nanoparticles appear
(994, 702, 284, and 146 cm−1 ).(97) 9 OPERANDO RAMAN STUDIES AND
RAMAN SPECTRA DURING REACTION
8.2 Temperature-programmed Desorption and WITH SIMULTANEOUS
Temperature-programmed Surface ACTIVITY/SELECTIVITY
Reaction-Raman DETERMINATION
Temperature-programmed Raman analyses are
particularly informative during TPSR (temperature- 9.1 VSbO4 Transformations During Propane
programmed surface reaction) of probe molecules. Ammoxidation Reaction: A Temperature-resolved
The Raman spectrum of dehydrated supported Operando Raman-Gas Chromatography Study
CrO3 /SiO2 gives rise to two bands at 982 and Bulk mixed metal oxides are efficient catalysts for
1011 cm−1(205) . It is not clear if these bands are the propane ammoxidation to acrylonitrile (Equation 5):
symmetric and asymmetric vibrations from one surface
CrOx species or from two different surface CrOx species. C3 H6 + O2 + NH3 → C3 H3 N + H2 O (5)
To address this issue, time-resolved reduction experi-
ments in H2 were undertaken at 500 ° C. The reduction Bulk antimony–vanadium mixed metal oxides are
treatment clearly showed that these vibrations origi- typical catalysts for this reaction(248 – 253,254 – 256) ; Raman
nated from two different surface CrOx species since spectra of fresh and used catalysts show significant

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
18 RAMAN SPECTROSCOPY

Cr–O-Cr
Under methanol Cr = O
Zr-OCH3

Formate Calcined Calcined

1596
5 min 5 min

35 min
35 min

132 min
132 min
Intensity (a.u.)

100 °C
100 °C

150 °C
150 °C

200 °C
200 °C

250 °C
250 °C

300 °C 300 °C

350 °C Formate
350 °C
Cr-OCH3

3100 3000 2900 2800 1600 1200 800


−1)
Raman shift (cm
Figure 12 Raman spectra of zirconia-supported chromate species during methanol TPSR and reoxidation. (Reproduced from
Ref. 247.  Elsevier, 2007.)

changes: surface vanadium oxide and surface anti- active phase. A similar investigation of the VNbO4 rutile
mony oxide species on alumina react to form rutile- phase under various environments showed that this phase
VSbO4 phases. Rutile vanadium antimonates exhibit formed and was stable in inert or reducing environments,
a Raman(257) and an IR(258) band at 880 cm−1 , char- but it segregated into the individual oxides in oxidizing
acteristic of the M–O–M vibration next to a cation atmospheres.(259)
vacancy.(257) The formation of VSbO4 phases and
their role in the ammoxidation reaction were investi-
9.2 Metal Molybdate Bulk-phase Transitions During
gated by operando Raman-GC(69) on alumina-supported
Reaction, a Time-resolved Operando Resonance
nanoscaled antimony–vanadium oxide catalysts. No
Raman-Gas Chromatography Study
appreciable differences were found in conversion and
selectivity in comparison to those recorded with a Mestl et al. reported detailed resonance-enhanced
conventional fixed-bed microreactor with a commer- Raman studies of mixed oxides possessing Mo5 +
cial bulk catalysts. The Raman mode of the surface ions.(43,44,140,141,240) RR enhancement is highly sensi-
vanadium oxide species near 1024 cm−1 tended to disap- tive to the addition of vanadium and tungsten oxides
pear during catalytic operation, and the shape of the on the formation of Mo5 O14 -type structures. This is
broad Raman band centered at 900 cm−1 also changed due to the intervalence charge transfer between penta-
(Figure 13). As propane ammoxidation became measur- coordinated Mo5+ and hexa-coordinated Mo6+(44,260) .
able at increasing temperatures, a broad Raman band They compared the performance for propene conver-
became evident at about 800 cm−1 (VSbO4 ).(69) The sion to acrylic acid. The highest propene conversion
concomitant formation of VSbO4 and Sb2 O4 and the and the highest selectivity to partial oxidation prod-
change in its selectivity from ODH to ammoxidation ucts were found for the starting prereduced material
underlines the relevance of the rutile VSbO4 as the (MoO3−x ). Propene conversion and selectivity to partial

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 19

SbVO4

Sb2O4

V=O
29.2 f Reoxidized
4.6
22.0 480 °C 480 °C
18.1
10.9
*
4.9 e
1.4 420 °C
5.2 420 °C
8.1
4.3
400 °C d
2.6
Ammoxidation
Acrylonitrile 1.0
4.1 400 °C c
Acetonitrile 4.9 200 °C
Propylene 2.4
CO2 b
CO
200 °C
a Dehydrated

30 20 10 0 1100 900 700 500 300 100


Yield (%) Raman shift (cm−1)
Figure 13 Operando Raman VSbO4 ammoxidation. (Reproduced from Ref. 69.  Royal Society of Chemistry, 2002.)

oxidation products (Figure 14) decrease with progressive environmentally unfriendly process. Methanol-induced
reoxidation of MoO3−x . Thus, activity and selectivity of dispersion allows preparing bulk iron molybdate catalysts
MoO3−x catalysts are a function of the degree of reduc- out of its constituting oxides. The fresh precursor mixture
tion and imply that oxygen defects are necessary for exhibits Raman bands of a-MoO3 and iron oxide is
catalytic activity.(44,109) More recent studies revealed that not detected due to the much lower Raman cross
the surfaces of such bulk mixed metal molybdates are section of α-Fe2 O3 . After exposure to the methanol
surface enriched with surface-MoOx sites that are the oxidation environment, new Raman bands evidence
catalytic active sites.(261,262) Future studies will have to the ready formation of bulk Fe2 (MoO4 )3 from the
further focus on the surface aspects of such bulk mixed mixture of the individual metal oxides.(138,139) Reaction-
metal molybdates. induced redispersion has also successfully been applied to
redisperse MoO3 in industrially spent bulk Fe2 (MoO4 )3
9.3 Raman Spectra of Redispersion/Regeneration catalysts that enhances both methanol activity and HCHO
During Reaction selectivity.(138,139) A section below illustrates the potential
Metal oxides can either disperse or aggregate under of time- and space-resolution during methanol oxidation
reaction conditions(22) and recent investigations have by a bulk iron molybdate catalyst.
demonstrated how to take advantage of methanol-
vapor-induced changes of the dispersion of the metal
oxide active phase to prepare industrial catalysts.(138,139)
10 CHEMICAL IMAGE MICROSCOPY
Reaction-induced dispersion may be used as an elegant
simple alternative to conventional preparation methods
for supported metal oxides(263) and is an application The ability to connect a Raman system to a microscope
of the solid–solid wetting process(264) for catalyst permits chemical imaging of the samples. The Raman
preparation. For example, the preparation of bulk iron mapping records complete Raman spectra in the selected
molybdate catalysts for methanol oxidation requires the spots. Imaging applications span the range of semi-
coprecipitation of an aqueous solution of iron trichloride conductors, microelectromechanical systems, fuel cells,
and ammonium heptamolybdate. This is a laborious and polymers, pharmaceuticals, meteorites, and biomedical

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
20 RAMAN SPECTROSCOPY

100 3.0

80

Raman intensity ratio I823 / I0823


2.5
Intensity (a.u.)

Selectivity (%)
60
2.0

10 h 40
1.5

5h 20

1.0
1h
0
1000 800 600 0 2 4 6 8 10
(a) Raman shift (cm−1) (b) Time on stream (h)

3.0
Propene conversion (%)

2.5

2.0

1.5

1.0

0.5

0.8 1.2 1.6 2.0 2.4


(c) Raman intensity ratio I823 / I0823

Figure 14 Catalytic tests of pure reoxidized MoO3 catalyst during propene conversion to acrylic acid: (a) (left axis) reaction
selectivity to acetic acid (open circles) and selectivity to CO2 (open upward triangles), and the intensity (I ) of the Raman band of
MoO3 at 823 cm−1 referenced to its starting intensity I0 (filled squares, right axis) as function of time on stream during propene
oxidation. (b) Propene conversion during the in situ Raman experiment as function of the ratio of the intensity of the prominent
MoO3 band at 823 cm−1 and its initial intensity (i.e. with time onstream). (Adapted, reproduced with permission, Bañares and Mestl,
Adv. Catal. 52 (2009) 43.)

materials.(265) Raman imaging also finds its applica- and molybdenum species in alumina pellets.(269) Wachs
tion in heterogeneous catalysis, like in the study of and Briand(138,139) recently employed Raman mapping
the dispersion of molybdenum oxide on silica and on to determine the distribution of crystalline MoO3
alumina supports.(264) In addition, novel microreactors in spent commercial bulk iron molybdate methanol
and membrane microreactors are becoming increasingly oxidation catalysts in pellet form (Figure 15). Bulk
popular owing to their unique properties.(266 – 268) iron molybdate catalysts contain both MoO3 and
Fe2 (MoO4 )3 crystalline phases. The catalyst pellet cross
section in Figure 16 reveals that MoO3 (indicated in
10.1 Raman Microspectroscopy, Scanning Near-field
yellow-green) was reasonably well dispersed over the
Optical Microscopy and Tip-enhanced Raman
Fe2 (MoO4 )3 phase (indicated in purple) in the fresh
Spectroscopy
catalyst. The Raman mapping of the spent catalyst,
Raman imaging is finding increasing applications in however, showed that MoO3 was absent from the outer
heterogeneous catalysis, for example, in the investigation regions of the pellets, but was present in the pellet
of the spreading of molybdenum oxide on silica and interior. This nonuniform distribution was a result of
on alumina supports(264) and the distribution of cobalt the volatilization of Mo–OCH3 species and of methanol

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 21

Spent catalyst

Fe2(MoO4)3
MoO3
Fe2(MoO4)3
MoO3

700 750 800 850 900


700 750 800 850 900
Raman shift (cm−1)
Raman shift (cm−1)
0 500
Length X (µm)

Fresh catalyst

MoO3
Fe2(MoO4)3

700 750 800 850 900


Raman shift (cm−1)

0 500
Length X (µm)
Figure 15 Raman imaging reveals MoO3 leaching out from the external surface of a catalyst pellet. (Source: I. E. Wachs).

transport limitations associated with its rapid oxidation spatial resolution and can be applied under the conditions
to formaldehyde in the outer region of the pellet. This of heterogeneous catalysis. The combined SNOM-Raman
insight led to a novel method for in situ regeneration of experiments simultaneously provide information about
bulk iron molybdate catalysts in commercial plants by sample topography and structure.
methanol treatments for 15–30 min. In the absence of Raman microspectroscopy has been recently combined
reactive O2 , methanol diffuses into the interior regions of with SNOM.(273 – 276,278,280 – 288) In an SNOM experiment,
the catalyst pellets and redistributes MoO2 to the exterior a hollow, metal-plated light-transmitting fiber tip is
via transport as Mo–OCH3 species. placed at such a distance to the surface that attractive
Spatially resolved Raman spectroscopy has also van der Waals forces can be used to control the tip
provided insights into the physicochemical processes motion in the z-direction, as in noncontact AFM. This
that determine the distribution of the H2 PMoO11 CoO40 obviously provides topographic information about the
active phase in alumina pellets.(269) Molybdenum and sample surface. In addition, the fiber tip has a diameter
cobalt complexes were found to diffuse through the of about 50 nm; hence, the laser light transmitted through
pore structure of the alumina pellets at different the tip excites the Raman spectrum of sample areas
rates: the transport of cobalt complexes was fast, smaller than l/10 nm. Thus, combined SNOM–Raman
whereas molybdenum required several hours to reach experiments provide simultaneous molecular (Raman)
even distribution. Spatially resolved Raman monitoring and topographic information. Because of the small
provides information about how preparation conditions Raman scattering cross sections of adsorbed molecules,
affect the distribution of molybdenum ions.(269) SNOM–Raman experiments are often performed with
Novel microreactors and membrane microreactors are SERS-active metal surfaces.(289 – 295)
also becoming increasingly popular because of their An effort to implement Raman SNOM resulted in the
unique mass and heat transfer properties.(266,267,270,271) development of TERS; in this case, the optical fiber is
Catalysts in such microfabricated reactors may be investi- replaced by an apertureless tip made of metal, which
gated by Raman imaging methods.(272) Furthermore, the enhances the Raman signal. This interaction triggers the
combination of scanning near-field optical microscopy Raman signal while provides topographic information.
(SNOM) with Raman spectroscopy(273 – 288) yields high The great advantage is that TERS may connect molecular

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
22 RAMAN SPECTROSCOPY

Front Middle Back

Nitrates
400 10

Nitrites

Rich
300

A /mAU
t (s)

5
200

Lean
100

0 0
1800 1400 1000 1800 1400 1000 1800 1400 1000
(a) (b) n(cm−1) (c)

1
400

Rich
300
0.5
t (s)

200

Lean
100

0 0
1050 1000 950 900 1050 1000 950 900 1050 1000 950 900
(d) (e) Raman shift (cm−1) (f)
Figure 16 Time-resolved DRIFT (a)–(c) and Raman (d)–(f) spectra maps during NSR at the front, middle, and back positions
of the catalyst bed. (Reproduced from Ref. 299.  John Wiley & Sons, Ltd, 2008.)

Raman information and topographic information. A the stored NOx is then released and reduced to N2
comprehensive special issue was published in Journal over the noble metal during fuel-rich conditions; the
of Raman Spectroscopy.(296) There is a great potential in Ba component is subsequently regenerated for a new
applying such an approach to catalytic materials under in cycle. A variety of Ba species, such as nitrite, nitrate,
situ conditions. carbonate, oxide, peroxide, and hydroxide, are involved
in the processes occurring during lean–rich cycles,
10.2 Operando Time- and Space-resolved Raman often appearing as overlapping signals. A combined use
of both diffuse reflectance infrared Fourier transform
Recent progress on in situ Raman systems allows spectroscopy (DRIFTS) and Raman at different catalyst-
for advanced combination of time- and space-resolved bed positions tackles the complex catalytic processes.(299)
analyses. In the following paragraphs, we comment on Figure 16(a)–(f) shows remarkable spectral differences
two representative studies that illustrate the potential for the different positions along the catalyst bed. Nitrites
to understand catalyst operation, deactivation, and and nitrates formation is significantly delayed along
regeneration. It is envisioned that progress will be catalyst bed. In the DRIFT spectra (Figure 16(a)–(c)), the
made to implement SNOM, and particularly TERS to most striking difference that occurs during lean periods,
operando conditions, or at least under in situ conditions. besides the band intensities, is the formation of nitrites,
A comprehensive overview on Raman imaging during which was detected immediately at the front position, but
reaction has been published recently.(297,298) delayed by 70 and 120 s at the middle and back positions,
Time- and space-resolved operando Raman studies respectively. Conversely, nitrates were formed from the
during nitrogen oxides storage/reduction (NSR) reaction beginning of the lean periods, independently of the bed
NSR is increasingly important due to its NOx reduction position, because of more complete NO oxidation to
capability in oxygen-rich conditions. NSR utilizes periodic NO2 over Pt toward the end of the bed (back position).
switching between oxidative atmosphere and reductive The temporal profiles of surface nitrate bands at 1028
atmosphere (fuel-rich); during fuel-lean periods, NO is and 1570 cm−1 were similar to that at 1260 cm−1 , while
oxidized to NO2 on Pt sites, and stored as barium nitrates; the band position of bulk ionic nitrate, visible at the

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 23

front position (Figure 16a), gradually shifted from 1750 development in the use of advanced multitechnique
to 1770 cm−1 during lean periods. The band at 1750 cm−1 operando analyses.
is not due to bulk Ba(NO2 )2 , as confirmed by Raman
spectroscopy, but rather caused by the formation of bulk
nitrates. The chemical changes, which occur during the ABBREVIATIONS AND ACRONYMS
NOx storage process, are likely accompanied by gradual
structural reordering, that is, changes in the crystallinity AFM Atomic Force Microscopy
of the solid during nitrate penetration from the surface CCD Charge-Coupled Device
into the bulk of Ba components. Furthermore, during DFT Density Functional Theory
rich periods, a remarkable difference in the reduction DRIFTS Diffuse Reflectance Infrared Fourier
behaviors of nitrites and nitrates was detected. Nitrites Transform Spectroscopy
were first reduced or desorbed from the surface and then DRS Diffuse Reflectance Spectroscopy
the reduction of nitrates followed, as clearly indicated ESR Electron Spin Resonance
by a significant delay of the corresponding signals at EXAFS Extended X-Ray Absorption Fine
the middle and back positions. The aforementioned large Structure
amount of NO, released from the catalyst at the beginning FFR Far-from Resonance
of rich periods, is likely related to the disappearance of FT Fourier Transform
surface nitrites, thus, suggesting decomposition of the GC Gas Chromatography
nitrites and consequent release of NO into the gas phase. IR Infrared
Moreover, the Raman spectra in Figure 16(d)–(f) clearly IEP Isoelectric Point
indicate the difference in the amount of formed bulk MS Mass Spectroscopy
Ba(NO3 )2 . At the front position, the amount increased NIR near Infrared
drastically. However, the increase was considerably less NSR Nitrogen Oxides Storage/reduction
prominent toward the back position. A similar tendency NMR Nuclear Magnetic Resonance
is also evident in the DRIFTS spectra in Figure 16(a)–(c), ODH Oxidative Dehydrogenation
which show much less intense surface nitrite and nitrate PZC Point of Zero Charge
bands toward the back position. Remarkably, 90% of RR Resonance Raman
the NOx remained stored at the end of lean periods, SNOM Scanning near-Field Optical Microscopy
but the majority of the NOx was stored at the front SES Single-Electronic-State
of the catalyst bed, utilizing the bulk Ba component. SERS Surface-Enhanced Raman Spectroscopy
The combined DRIFTS–Raman approach elucidates TPO Temperature-Programmed Oxidation
the position-dependent bulk utilization of the storage TPR Temperature-Programmed Reduction
component, which greatly affects the efficiency of the TPSR Temperature-Programmed Surface
NOx storage process. Reaction
TERS Tip-Enhanced Raman Spectroscopy
UV–vis Ultraviolet–Visible
XANES X-Ray Absorption near Edge Structure
11 MULTITECHNIQUES XRD X-Ray Diffraction

The use of several spectroscopic techniques simultane-


ously to Raman measurements allows obtaining comple- RELATED ARTICLES
mentary simultaneous information from the sample
during the treatment, synthesis, and catalytic process. Dispersive Raman Spectroscopy, Current Instrumental
The previous section has illustrated the combined use Designs
of Raman and DRIFTS on nitrogen storage-reduction Fourier Transform Raman Instrumentation
catalysts.(299) Several authors report combined use of Infrared and Raman Spectroscopy and Imaging in
spectroscopies in a simultaneous fashion, such as Raman Coatings Analysis
and UV–vis(300) or three techniques: Raman, EPR, Infrared and Raman Spectroscopy in Analysis of Surfaces
and UV–vis(301) to study catalyst activity, deactivation, Raman Microscopy and Imaging
and regeneration. The use of multiple spectroscopic Raman Scattering, Fundamentals
techniques in a simultaneous fashion has also been Raman Spectroscopy in Process Analysis
used to monitor/study the synthesis of catalysts.(302) It is Raman Spectroscopy in Analysis of Biomolecules
envisioned that in the near future there will be major Raman Spectroscopy: Introduction

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
24 RAMAN SPECTROSCOPY

REFERENCES 28. W.N. Delgass, G.L. Haller, R. Kellerman, J.H. Lunsford,


Spectroscopy in Heterogeneous Catalysis, Academic
1. J.M. Thomas, Characterization of Catalysts, eds. Press, New York, 55, 1979.
J.M. Thomas, R.M. Lambert, John Wiley and Sons, New
29. M. Delhaye, P. Dhamelincourt, J. Raman Spectrosc., 3,
York, 1980.
33 (1997).
2. H. Topsøe, Stud. Surf. Sci. Catal., 130, 1–21 (2000). 30. L. Dixit, D.L. Gerrard, H.J. Bowley, Appl. Spectrosc.
3. B.M. Weckhuysen, Chem. Commun., 97 (2002) (Feature Rev., 22, 189 (1986).
Article). 31. T.A. Egerton, A.H. Hardin, Catal. Rev.-Sci. Eng., 11, 1
4. R. Burch (ed.), special issue ‘In situ Methods in Catalysis’, (1975).
Catal. Today, 9(1/2), (1991). 32. M. Fleischmann, P.J. Hendra, A.J. McQuillan, R.L. Paul,
5. B.S. Clausen, H. Topsøe, R. Frahm, Adv. Catal., 42, 315 E.S. Reid, J. Raman Spectrosc., 4, 269 (1976).
(1998). 33. E. Garbowski, G. Coudurier, Catalyst Characterization.
6. J.A. Dumesic, H. Topsøe, Adv. Catal., 26, 121 (1977). Physical Techniques for Solid Materials, eds. B. Imelik,
J.C. Vedrine, Plenum Press, New York, 45, 1994.
7. M. Hunger, J. Weitkamp, Angew. Int. Ed., 40, 2954
34. J.G. Grasselli, M.A.S. Hazle, L.E. Wolfram, Molecular
(2001).
Spectroscopy, ed. A.R. West, Heyden and Son, London,
8. J.W. Niemantsverdriet, Spectroscopy in Catalysis, VCH, 200, 1977.
1993.
35. W. Kiefer, H.J. Bernstein, Appl. Spec, 25, 609 (1971).
9. G.A. Somorjai, Cattech, 3, 84 (1999).
36. W. Kiefer, H.J. Bernstein, Appl. Spectrosc., 25, 500
10. J.M. Thomas, G.A. Somorjai, Topics Catal., 8(1/2), (1971).
preface (1999). 37. H. Knözinger, G. Mestl, Topics Catal., 8, 45 (1999).
11. G.A. Somorjai, G. Rupprechter, J. Phys. Chem. B, 103, 38. H. Knözinger, Characterization of Heterogeneous Cata-
1623 (1999). lysts Studied by Particle Beams, eds. H.H. Brongersma,
12. F. Zaera, Prog. Surf. Sci., 69, 1 (2001). R.A. van Santen, Plenum Press, New York, 167, 1991.
13. F. Zaera, Surf. Sci., 55, 947 (2002). 39. H. Knözinger, Catal. Today, 32, 71 (1996).

14. Y. Gauthier, R. Baudoing-Savois, K. Heinz, 40. S. Kuba, H. Knözinger, J. Raman Spectrosc., 33, 325
H. Landskron, Surf. Sci., 251, 493 (1991). (2002).

15. H.E. Shih, F. Jonas, P.M. Marcus, Phys. Rev. Lett., 46, 41. M. Mehicic, J.G. Grasselli, Analytical Raman Spec-
731 (1981). troscopy, eds. J.G. Grasselli, B.J. Bulkin, Wiley,
Chichester, 325, 1991.
16. G.A. Somorjai, Chem. Rev., 96, 1223 (1996).
42. G. Mestl, H. Knözinger, Handbook of Heterogeneous
17. I.E. Wachs, Surf. Sci., 544, 1 (2003). Catalysis, eds. G. Ertl, H. Knözinger, J. Weitkamp,
18. I.E. Wachs, Catal. Commun, 4, 567 (2003). Wiley-VCh, Weinheim, 539, Vol. 2, 1997.

19. I.E. Wachs, CATTECH., 7, 142 (2003). 43. G. Mestl, J. Mol. Catal. A, 158, 45 (2000).

20. M.A. Bañares, Catal. Today, 100, 71 (2005). 44. G. Mestl, J. Raman Spectrosc., 33, 333 (2002).
45. G. Mestl, T.K.K. Srinivasan, Catal. Rev.- Sci. Eng., 40,
21. I.E. Wachs, Catal. Today, 100, 79 (2005).
451 (1998).
22. M.A. Bañares, G. Mestl, Adv. Catal., 52, 43 (2009).
46. G. Mestl, H. Knözinger, J.H. Lunsford, Ber. Bunsen Ges.
23. M.A. Bañares, I.E. Wachs, J. Raman Spectrosc., 33, 359 Phys. Chem., 97, 319 (1993).
(2002). 47. B.A. Morrow, Vibrational Spectroscopy of Adsorbed
24. M.A. Bañares, In situ Characterization of Catalytic Species, ACS. Symp. Series, eds. A.T. Bell, M.L. Hair,
Materials, ed. B.M. Weckhuysen, American Scientific ACS, Washington, DC, 119, Vol. 137, 1981.
Publishers, 2004. 48. T.T. Nguyen, J. Singapore Natl. Acad. Sci., 10-12, 84
25. J.R. Bartlett, R.P. Cooney, Spectroscopy of Inorganic- (1983).
Based Materials, Advances in Spectroscopy, eds. 49. E. Payen, S. Kasztelan, Trends Phys. Chem., 4, 363 (1994).
R.J.H. Clark, R.E. Hester, Wiley, Chichester, 187, Vol.
50. E. Payen, J. Barbillat, J. Grimblot, J.P. Bonnelle,
14, 1987.
Spectrosc. Lett., 11, 997 (1978).
26. M.A. Chesters, N. Sheppard, Chem. Brit., 17, 521 (1981). 51. K. Segawa, I.E. Wachs, Characterization of Catalytic
27. R.P. Cooney, G. Curthoys, T.T. Nguyen, Adv. Catal., 24, Materials, ed. I.E. Wachs, Butterworth-Heinemann,
293 (1975). Boston, 1992.

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 25

52. N. Sheppard, J. Erkelens, Appl. Spectrosc., 38, 471 (1984). 78. L.A. Nafie, Theory of Raman Scattering, Handbook of
53. P.C. Stair, Curr. Opin. Surf. Sci. Mat. Sci., 5, 365 (2001). Raman Spectroscopy, eds. I.R. Lewis, H.G.M. Edwards,
Marcel Dekker, New York, Chap.1, 2001.
54. J.M. Stencel, Raman Spectroscopy for Catalysis, Van
79. Y.T. Chua, P.C. Stair, J. Catal., 196, 66 (2000).
Nostrand, Reinhold, New York, 1990.
80. G. Xiong, C. Li, Z. Feng, J. Li, P. Ying, H. Li, Q. Xin,
55. T. Takenaka, Adv. Colloid. Interf. Sci., 11, 291 (1979).
Studs. Surf. Sci. Catal., 130, 341 (2000).
56. H. Tian, I.E. Wachs, L.E. Briand, J. Phys. Chem. B, 109,
81. D.C. Smith, C. Carabatos-Nédelec, Handbook of Raman
23491 (2005).
Spectroscopy, Chapter 9, Raman Spectroscopy Applied
57. I.E. Wachs, F.D. Hardcastle, S.S. Chan, Spectroscopy, 1, to Crystals: Phenomena and Principle, Concepts and
30 (1986). Conventions, eds. I.R. Lewis, H.G.M. Edwards, Marcel
58. I.E. Wachs, Catal. Today, 27, 437 (1996). Decker, New York, 401, 2001.
59. I.E. Wachs, Topics Catal., 8, 57 (1999). 82. F.A. Cotton, R.M. Wing, Inorg. Chem., 4, 867 (1965).

60. I.E. Wachs, Raman Scattering in Materials Science, eds. 83. F.D. Hardcastle, I.E. Wachs, J. Raman Spectrosc., 27, 683
W.H. Weber, R. Merlin, Springer, Berlin, 271, 2000. (1990).

61. I.E. Wachs, Handbook of Raman Spectroscopy, ed. 84. F.D. Hardcastle, I.E. Wachs, J. Phys. Chem., 95, 5031
I.R. Lewis, Marcel Dekker, 2002. (1991).

62. I.E. Wachs, F.D. Hardcastle, in Catalysis, The Royal 85. F.D. Hardcastle, I.E. Wachs, Solid State Ionics, 45, 201
Society of Chemistry, Cambridge, 102, Vol. 10, 1993. (1991).
86. F.D. Hardcastle, I.E. Wachs, J. Solid State Chem., 97, 319
63. M.J. Weaver, J. Raman Spectrosc., 33, 309 (2002).
(1992).
64. W.H. Weber, Springer Ser. Mater. Sci., 42, 233 (2000).
87. F.D. Hardcastle, I.E. Wachs, J. Raman Spectrosc., 26, 397
65. W.H. Weber, Mater. Sci., 42, 233 (2000). (1995).
66. M.A. Bañares, M.O. Guerrero-Pérez, G. Garcı́a-Cortéz, 88. A. Rulmont, R. Cahay, M. Liegeois-Duyckaerts,
J.L.G. Fierro, J. Mater. Chem., 11, 3337 (2002). P. Tarte, Eur. J. Solid. State Inorg. Chem., 28, 207 (1991).
67. B.S. Clausen, G. Steffensen, B. Fabius, J. Villadsen, 89. B.M. Weckhuysen, I.E. Wachs, Journal of the Chemical
R. Feidenhans’l, H. Topsøe, J. Catal., 132, 524 (1991). Society - Faraday Transactions, 92, 1969 (1996).
68. G. Garcı́a-Cortéz, M.A. Bañares, J. Catal., 209, 197 90. I.D. Brown, K.K. Wu, Acta Cryst., B32, 1957 (1976).
(2002).
91. J.R. Ferraro, K. Nakamoto, Introductory Raman
69. M.O. Guerrero-Pérez, M.A. Bañares, Chem. Commun., Spectroscopy, Academic Press, New York, 1994.
1292 (2002). 92. I.R., Lewis, H.G.M., Edwards (eds.), Handbook of
70. M.O. Guerrero-Pérez, M.A. Bañares, Catal. Today, 96, Raman Spectroscopy, Marcel Dekker Inc., New York,
265 (2004). 2001.
71. P. Rybaczyk, H. Berndt, J. Radnik, M.M. Pohl, 93. H.O. Hamaguchi, Appl. Spectrosc. Rev., 24, 137 (1988).
O. Buyevskaya, M. Baerns, A. Brückner, J. Catal., 202, 94. J.B. Slater, J.M. Tedesco, R.C. Fairchild, I.R. Lewis,
45 (2001). ‘Chapter 3: Raman Spectroscopy and its adaptation to
72. D. Tibiletti, A. Goguet, F.C. Meunier, J.P. Breen, the industrial Environment’, in Handbook of Raman
R. Burch, Chem. Commun., 1636 (2004). Spectroscopy, eds. I.R. Lewis, H.G.M. Edwards, Marcel-
Dekker, New York, 2001.
73. I.E. Wachs, F.D. Hardcastle, Catalysis – A Specialist
Periodical Report, Royal Society of Chemistry, 95. C.L. Pieck, M.A. Bañares, M.A. Vicente, J.L.G. Fierro,
Cambridge, UK, 102 Vol. 10, 1993. Chem. Mater., 13, 1174 (2001).
74. B.M. Weckhuysen, Phys. Chem. Chem. Phys., 5, 4351 96. G.I.N. Waterhouse, G.A. Bowmaker, J.B. Metson, Appl.
(2003). Surf. Sci., 214, 36 (2003).
75. I.E. Wachs, ‘Chapter 20. Raman Spectroscopy of 97. M.A. Bañares, J.H. Cardoso, F. Agulló-Rueda, J.M.
Catalysts’, in Handbook of Raman Spectroscopy, eds. Correa-Bueno, J.L.G. Fierro, Catal. Lett., 64, 191 (2000).
I.R. Lewis, H.G.M. Edwards, Marcel-Dekker, New York, 98. J.M. Kanervo, M.E. Harlin, A.O.I. Krause, M.A.
2001. Bañares, Catal. Today, 78, 171 (2003).
76. M. Delhaye, P. Dhamelincourt, J. Raman Spectrosc., 3, 99. J.E. Herrera, D.E. Resasco, Chem. Phys. Lett., 376, 302
33 (1973). (2003).
77. Y.T. Chua, P.C. Stair, I.E. Wachs, J. Phys. Chem. B, 105, 100. L.J. Burcham, G. Deo, X. Gao, I.E. Wachs, Topics Catal.,
8600 (2001). 11/12, 85 (2000).

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
26 RAMAN SPECTROSCOPY

101. J.G. Grasselli, M.K. Snavely, J.J. Bulkin, Chemical 127. S. Xie, M.P. Rosynek, J.H. Lunsford, Appl. Spectrosc., 53,
Applications of Raman Spectroscopy, Wiley Interscience, 1183 (1999).
1981. 128. G.L. Haller, J.H. Lunsford, J. Laane, J. Phys. Chem. 100,
102. F.J. Bergin, Spectrochim. Acta A, 46, 153 (1990). 551 (1996).
103. R.G. Messerschmidt, D.B. Chase, Appl. Spectrosc., 43, 11 129. G. Mestl, M.P. Rosynek, J.H. Lunsford, J. Phys. Chem.
(1989). B, 102, 154 (1998).
104. G.J. Rosasco, Advances in Infrared and Raman 130. S. Xie, G. Mestl, M.P. Rosynek, J.H. Lunsford, J. Am.
Spectroscopy, eds. R. Clark, R.E. Hester, Heyden and Chem. Soc., 119, 10186 (1997).
Son, London, 223, Vol. 7, 1980. 131. J.A. Koningstein, B.F. Gachter, J. Opt. Sci. Am., 63, 882
105. B. Schrader, Fresenius J. Anal. Chem., 337, 824 (1990). (1973).
106. A.J. Sommer, J.E. Katon, Appl. Spectrosc., 45, 527 (1991). 132. T.P. Snyder, C.G. Hill, Jr J. Catal. 132, 536 (1991).
107. P.J. Treado, M.D. Morris, Appl. Spectrosc. Rev., 29, 1 133. N. Zimmerer, W. Kiefer, Appl. Spectrosc., 28, 279 (1974).
(1994). 134. L.J. Burcham, I.E. Wachs, Catal. Today, 49, 467 (1999).
108. E. Payen, M.C. Dhamelincourt, P. Dhamelincourt, 135. M.A. Bañares, S.J. Khatib, Catal. Today, 96, 251 (2004).
J. Grimblot, J.P. Bonelle, Appl. Spectrosc., 36, 30 (1980).
136. M.V. Martínez-Huerta, G. Deo, J.L.G. Fierro, M.A.
109. M. Dieterle, G. Mestl, J. Jäger, Y. Uchida, R. Schlögl, Bañares, J. Phys. Chem. C, 112, 11441–11447 (2008).
J. Mol. Catal. A, 174, 169 (2001).
137. J.A. Wilson, C.G. Hill, J.A. Dumesic, J. Mol. Catal. A, 61,
110. S.I. Woo, C.G. Hill, J. Mol. Catal., 15, 309 (1982). 333 (1990).
111. S.I. Woo, C.G. Hill, J. Mol. Catal., 24, 165 (1984). 138. I.E. Wachs, L.E. Briand, US Patent 2002/0062048, 2002.
112. S.I. Woo, C.G. Hill, J. Mol. Catal., 29, 231 (1985). 139. I.E. Wachs, L.E. Briand WO Patent WO 2002/022539,
113. F.B. Abdelouahab, R. Olier, N. Guilhaume, F. Lefebvre, 2002.
J.C. Volta, J. Catal., 134, 151 (1992). 140. M. Dieterle, PhD Thesis, Technical University, Berlin,
114. F.R. Brown, L.E. Makovsky, H.K. Rhee, Appl. Spec- 2001.
trosc., 31, 563 (1977). 141. O. Ovsitser, Y. Uchida, G. Mestl, G. Weinberg,
115. F.R. Brown, L.E. Makovsky, K.H. Rhee, J. Catal., 50, 162 A. Blume, J. Jäger, M. Dieterle, H. Hibst, R. Schlögl,
(1977). J. Mol. Catal. A, 185, 291 (2002).
116. F.R. Brown, L.E. Makovsky, K.H. Rhee, J. Catal., 50, 385 142. M.A. Bañares, M.V. Martínez-Huerta, X. Gao, J.L.G.
(1977). Fierro, I.E. Wachs, ‘Metal Oxide Catalysts: Active Sites,
117. S.S. Chan, A.T. Bell, J. Catal., 89, 433 (1984). Intermediates and Reaction Mechanisms’, Symposium,
220th ACS National Meeting, Washington, DC, 2000.
118. C.P. Cheng, J.D. Ludowise, G.L. Schrader, Appl.
Spectrosc., 34, 146 (1980). 143. A.A. Tolia, C.T. Williams, C.G. Takoudis, M.J. Weaver,
J. Phys. Chem., 99, 4599 (1995).
119. J.H. Lunsford, X. Yang, K. Haller, J. Laane, G. Mestl,
H. Knözinger, J. Phys. Chem., 97, 13810 (1993). 144. C.T. Williams, A.A. Tolia, M.J. Weaver, C.G. Takoudis,
Chem. Eng. Sci., 51, 1673 (1996).
120. G. Mestl, M.P. Rosynek, J.H. Lunsford, J. Phys. Chem.
B, 101, 9329 (1997). 145. J.M. Stencel, L.E. Makovsky, T.A. Sarkus, J. De Vries,
R. Thomas, J.A. Moulijn, J. Catal., 90, 314 (1984).
121. G. Mestl, M.P. Rosynek, J.J. Lunsford, J. Phys. Chem.,
101, 9321 (1997). 146. G.L. Schrader, C.G. Hill, Rev. Sci. Instrum., 46, 1335
(1975).
122. J.C. Vèdrine, E.G. Derouane, Combinatorial Catalysis
and High Throughput Catalyst Design and Testing, 147. S.S. Chan, I.E. Wachs, L.L. Murrell, L. Wang, K. Hall,
ed. E.G. Derouane, Kluwer Academic Publishers, New J. Phys. Chem., 88, 5831 (1984).
York, 2000. 148. H. Hu, I.E. Wachs, S.R. Bare, J. Phys. Chem., 99, 10897
123. A.K. Covington, J.M. Thain, Appl. Spectrosc., 29, 386 (1995).
(1975). 149. P.L. Villa, F. Trifirò, I. Pasquon, React. Kinet. Catal. Lett.,
124. V.V. Guliants, J.B. Benziger, S. Sundaresan, N. Yao, 1, 341 (1974).
I.E. Wachs, Catal. Lett., 32, 379 (1995). 150. H. Knözinger, H. Jeziorowski, J. Phys. Chem., 82, 2002
125. Q. Sun, J.-M. Jehng, H. Hu, R.G. Herman, I.E. Wachs, (1978).
K. Klier, J. Catal., 165, 101 (1997). 151. G.L. Schrader, C.P. Cheng, J. Catal., 80, 369 (1983).
126. M.A. Bañares, H. Hu, I.E. Wachs, J. Catal., 150, 407 152. R.J. Thomas, J.A. Moulijn, F.P.J.M. Kerkhof, Recl.
(1994). Trav.Chim. Pays-Bas, 96, M114 (1977).

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 27

153. L. Wang, K.W. Hall, J. Catal., 82, 177 (1983). 180. W.P. Griffith, J. Chem. Soc. A, 286 (1970).
154. F. Roozeboom, J. Medema, P.J. Gellings, Z. Phys. Chem., 181. W.P. Griffith, P.J.B. Lesniak, J. Chem Soc. A, 1066 (1969).
111, 215 (1978). 182. W.P. Griffith, T.D. Wickins, J. Chem. Soc. A, 675 (1967).
155. F.P.J.M. Kerkhof, J.A. Moulijn, R.J. Thomas, J. Catal., 183. S. Himeno, H. Niiya, T. Ueda, Bull. Chem. Soc. Jpn., 70,
56, 279 (1979). 631 (1997).
156. A. Iannibello, P.L. Villa, S. Marengo, Gazz. Chim. Ital., 184. W.D. Hunnius, Z. Naturforsch. B, 30, 63 (1975).
109, 521 (1979).
185. G. Johanson, L. Petterson, N. Ingri, Acta. Chem. Scand.
157. H. Jeziorowski, H. Knözinger, J. Phys. Chem., 83, 1166 A, 33, 305 (1979).
(1979).
186. K.Y. Ng, E. Gulari, Polyhedron, 3, 1001 (1984).
158. Y. Okamoto, T. Imanaka, J. Phys. Chem., 92, 7102 (1988).
187. G.L. Schrader, M.L. Batista, C.B. Bergman, Chem. Eng.
159. L. Wang, W.K. Hall, J. Catal., 66, 251 (1980). Commun., 12, 121 (1981).
160. C.C. Williams, J.G. Ekerdt, J.-M. Jehng, F.D. Hardcastle, 188. K.-H. Tytko, O. Glemser, Adv. Inorg. Chem. Radiochem.,
I.E. Wachs, J. Phys. Chem., 95, 8791 (1991). 19, 239 (1976).
161. R.J. Hunter, ‘Zetapotential in Colloid Science, Principle 189. K.-H. Tytko, J. Mehrnke, Z. Anorg. Allg. Chem., 503, 67
and Application’, in Colloid Science, eds. R.H. Ottewill, (1983).
R.L. Rowell, Academic Press, London, 233, 1988.
190. K.-H. Tytko, B. Schönfeld, Z. Naturforsch. B, 30, 63
162. G. Deo, I.E. Wachs, J. Phys. Chem., 95, 5889 (1991). (1975).
163. J. Gil-Llambias, A.M. Escudey, J.L.G. Fierro, A. López 191. K.-H. Tytko, B. Schönfeld, V. Cordis, O. Glemser, Z.
Agudo, J. Catal., 95, 520 (1985). Naturforsch. B, 30, 834 (1975).
164. F.D. Hardcastle, I.E. Wachs, J.A. Horsley, J. Mol. Catal., 192. K.-H. Tytko, G. Petridis, B. Schönfeld, Z. Naturforsch. B,
46, 15 (1988). 35, 45 (1980).
165. R.D. Roark, S.D. Kohler, J.G. Ekerdt, Catal. Lett., 16, 71 193. K.-H. Tytko, G. Baethe, E.R. Hirschfeld, J. Mehrnke,
(1992). D. Stellhorn, Z. Anorg. Allg. Chem., 503, 43 (1983).
166. R.D. Roark, S.D. Kohler, J.G. Ekerdt, D.S. Kim, 194. X. Carrier, J.F. Lambert, M. Che, J. Am. Chem. Soc., 119,
I.E. Wachs, Catal. Lett., 16, 77 (1992). 10137 (1997).
167. J.A. Horsley, I.E. Wachs, J.M. Brown, G.H. Via, 195. L. Le Bihan, P. Blanchard, M. Fournier, J. Grimblot,
F.D. Hardcastle, J. Phys. Chem., 91, 4014 (1987). E. Payen, J.Chem. Soc., Faraday Trans., 94, 937 (1998).
168. B.M. Weckhuysen, R.A. Schoonheydt, J.-M. Jehng, 196. K. Marcinkowska, A. Adnot, P.C. Roberge, S. Kalia-
I.E. Wachs, S.J. Cho, R. Ryoo, S. Kijlstra, E. Poels, guine, J. Phys.Chem., 90, 4773 (1986).
J. Chem. Soc. Faraday Trans., 91, 3245 (1995).
197. C. Rocchiccioli-Deltcheff, M. Amirouche, M. Che, J.M.
169. J.-M. Jehng, I.E. Wachs, J. Mol. Catal., 13, 9 (1992). Tatibouet, M. Fournier, J. Catal., 125, 2892 (1990).
170. J.-M. Jehng, I.E. Wachs, Chem. Mater., 3, 100 (1991). 198. J.M. Stencel, J.R. Diehl, J.R. D’Este, L.E. Makowsky,
171. M. Vuurman, I.E. Wachs, A.M. Hirt, J. Phys. Chem., 95, L. Rodrigo, K. Marcinkowska, A. Adnot, P.C. Roberge,
9928 (1991). S. Kaliaguine, J. Phys.Chem., 90, 4739 (1986).
172. I.E. Wachs, G. Deo, M. Vuurman, H. Hu, D.S. Kim, J.- 199. F.D. Hardcastle, I.E. Wachs, J. Phys. Chem., 46, 173
M. Jehng, J. Mol. Catal., 82, 443 (1993). (1988).
173. J.-M. Jehng, H. Hu, X. Gao, I.E. Wachs, Catal. Today, 28, 200. J.-M. Jehng, I.E. Wachs, J. Phys. Chem., 95, 7373 (1991).
335 (1996). 201. M.A. Vuurman, I.E. Wachs, J. Phys. Chem., 96, 5008
174. A. Christodoulakis, M. Machli, A.A. Lemonidou, (1992).
S. Boghosian, J. Catal., 222, 293 (2004). 202. X. Gao, J.L.G. Fierro, I.E. Wachs, Langmuir, 15, 3168
175. S. Xie, A.T. Bell, Catal. Lett., 70, 137 (2000). (1999).

176. X. Gao, S.R. Bare, B.M. Weckhuysen, I.E. Wachs, 203. X. Gao, I.E. Wachs, J. Catal., 192, 18 (2000).
J. Phys. Chem. B, 102, 10 842 (1998). 204. Y. Chen, J.L.G. Fierro, T. Tanaka, I.E. Wachs, J. Phys.
177. C. Baes, R.E. Mesmer, Hydrolysis of Cations, ed. Chem. B, 107, 5243 (2003).
F.A. Cotton, John Wiley & Sons, New York, 258, 1976. 205. E. Lee, I.E. Wachs, J. Phys. Chem. C, 111, 14410 (2007).
178. C.P. Cheng, G.L. Schrader, J. Catal., 60, 276 (1979). 206. E. Lee, I.E. Wachs, J. Phys. Chem. C, 112, 6487 (2008).
179. F. Gonzales-Vildres, W.P. Griffith, J. Chem Soc., Dalton 207. G. Deo, I.E. Wachs, J. Haber, Crit. Rev. Surf. Chem., 4,
Trans., 13, 1416 (1972). 141 (1994).

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
28 RAMAN SPECTROSCOPY

208. H.J. Freund, Catal. Today, 100, 3 (2005). 232. Z. Zhao, X. Gao, I.E. Wachs, J. Phys. Chem. B, 107, 6333
209. N. Magg, B. Immaraporn, J.B. Giorgi, T. Schröder, (2003).
M. Bäumer, J. Döbler, Z.L. Wu, E. Kondratenko, 233. C.G. Hill, J.H. Wilson, J. Mol. Catal., 63, 65 (1990).
M. Cherian, M. Baerns, P.C. Stair, J. Sauer, H.J. Freund, 234. C.L. Angell, J. Phys Chem., 77, 222 (1973).
J. Catal., 226, 88 (2004).
235. H. Barasnka, B. Czerwinska, A. Labudzinska, J. Mol.
210. Z. Wu, H.S. Kim, P.C. Stair, S. Rugmini, S.D. Jackson, Struct., 143, 485 (1986).
J. Phys. Chem. B, 109, 2793 (2005).
236. P.K. Dutta, M. Puri, J. Phys. Chem., 91, 4329 (1987).
211. S. Xie, E. Iglesia, A.T. Bell, J. Phys. Chem. B, 105, 5144
237. F. Roozeboom, H. Robson, S.S. Chan, Zeolites, 3, 321
(2001).
(1983).
212. H. Tian, E.I. Ross, I.E. Wachs, J. Phys. Chem. B, 110,
238. C. Brémard, D. Bougeard, Adv. Mater., 7, 10 (1995).
9593 (2006).
239. J.-M. Jehng, I.E. Wachs, J. Raman Spectrosc., 22, 83
213. E.I. Ross-Medgaarden, I.E. Wachs, J. Phys. Chem. C, 111,
(1991).
15089 (2007).
240. S. Knobl, G.A. Zenkovets, G.N. Kryukova, O. Ovsitser,
214. E. Diemann, T. Weber, A. Muller, J. Catal., 148, 288
D. Niemeyer, R. Schlögl, G. Mestl, J. Catal., 215, 177
(1994).
(2003).
215. X. Gao, I.E. Wachs, M.S. Wong, J.Y. Ying, J. Catal., 203,
241. V.V. Guliants, S. Holmes, J.B. Benziger, P. Heaney,
18 (2001).
D. Yates, I.E. Wachs, J. Mol. Catal. A, 172, 265 (2001).
216. T. Bredow, E. Aprà, M. Catti, G. Pacchioni, Surf. Sci.,
242. J.C. Volta, K. Bere, Y.J. Zhang, R. Olier, ACS
418, 150 (1998).
Symposium Series, 523, 217 (1992).
217. G.G. Cortéz, J.L.G. Fierro, M.A. Bañares, Catal. Today, 243. M.A. Carreón, V.V. Guliants, M.O. Guerrero-Pérez,
78, 219 (2003). M.A. Bañares, Microporous Mesoporous Mater., 71, 57
218. U. Diebold, Surf. Sci. Reports, 48, 53 (2003). (2004).
219. M.A. Vuurman, D.J. Stufkens, A. Oskam, G. Deo, 244. K. Angoni, J. Mater Sci., 33, 3693 (1998).
I.E. Wachs, J. Chem. Soc. Faraday Trans., 92, 3259 (1996). 245. A. Cuesta, P. Dhamelincourt, J. Laureyns, A. Martı́nez-
220. S.S. Chan, I.E. Wachs, J. Catal., 103, 224 (1987). Alonso, J.M.D. Tascón, Carbon, 32, 1523 (1994).
221. X. Gao, S.R. Bare, J.L.G. Fierro, I.E. Wachs, J. Phys. 246. N. Vidano, D.B. Fischbach, J. Am. Ceram. Soc., 61, 13
Chem. B., 103, 618 (1999). (1978).
222. F. Cavani, F. Trifirò, CHEMTECH, 24, 18 (1994). 247. S.T. Korhonen, M.A. Bañares, J.L.G. Fierro, A.O.I.
Krause, Catal. Today, 126, 235 (2007).
223. P.K. Dutta, D.C. Shieh, J. Phys. Chem., 90, 2331 (1986).
248. A. Andersson, S. Hansen, A. Wickman, Topics Catal., 15,
224. P.K. Dutta, D.C. Shieh, M. Puri, J. Phys. Chem., 91, 2332
103 (2001).
(1987).
249. A. Barbaro, S. Larrondo, S. Duhalde, N. Amadeo, Appl.
225. G.J. Hutchings, A. Desmartin-Chomel, R. Olier, J.C.
Catal. A, 193, 277 (2000).
Volta, Nature, 368, 41 (1994).
250. J.F. Brazdil, J.P. Bartek, US Patent 5854172, 1998.
226. I.E. Wachs, G. Deo, B.M. Weckhuysen, A. Andreini,
M.A. Vuurman, M. de Boer, M.D. Amiridis, J. Catal., 251. J.F. Brazdil, F.A.P. Kobarvkantei, J.P. Padolwski, JP
161, 211 (1996). Patent 11033399, 1999.
227. I.E. Wachs, J.-M. Jehng, G. Deo, B.M. Weckhuysen, 252. L.E. Briand, W.E. Farneth, I.E. Wachs, Catal. Today, 62,
V. Guliants, J.B. Benziger, Catal. Today, 32, 47 (1996). 219 (2000).

228. S. Damyanova, L.M. Gomez-Sainero, M.A. Bañares, 253. G. Centi, S. Perathoner, Appl. Catal. A, 124, 317 (1995).
J.L.G. Fierro, Chem. Mater., 12, 501 (2000). 254. V. Cortés-Corberán, V.V. Savkin, P. Ruiz, V.P.
229. N.N. Greenwood, E. Earnshaw, Chemistry of the Vislovskii, J. Mol. Catal. A, 158, 271 (2000).
Elements, Pergamon Press, Oxford, New York, Toronto, 255. A.T. Guttmann, R.K. Grasselli, F.J. Brazdil, US Patent,
Sidney, Paris, Frankfurt, 1984. 4746641, 4788173, and 4837233, 1988.
230. J.M. Tatibouët, M. Che, M. Amirouche, M. Fournier, 256. S. Larrondo, B. Irigoyen, G. Barenetti, N. Amadeo,
C. Rocchiccioli-Deltcheff, J. Chem. Soc., Chem. Appl. Catal. A, 250, 279 (2003).
Commun., 1260 (1988). 257. G. Xiong, V.S. Sullivan, P.C. Stair, G.W. Zajac, S.S. Trail,
231. J.M. Leroy, S.L. Peirs, G. Tridot, Comptes R. Acad. Sci., J.A. Kaduk, J.T. Golab, J.F. Brazdil, J. Catal., 230, 317
272, 218 (1971). (2005).

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
RAMAN SPECTROSCOPY OF CATALYSTS 29

258. A. Landa-Cánovas, J. Nilsson, S. Hansen, K. Stahl, 284. D.W. Pohl, W. Denk, M. Lanz, Appl. Phys. Lett., 44, 651
A. Andersson, J. Solid State Chem., 116, 369 (1995). (1984).
259. F. Cavani, N. Ballarini, M. Cimini, F. Trifirò, M.A. 285. Y. Tanaka, K. Fukuzawa, H. Kuwano, J. Appl. Phys., 83,
Bañares, M.O. Guerrero-Pérez, Catal. Today, 112, 12 3547 (1998).
(2006).
286. R. Toledo-Crow, P.C. Yang, Y. Chen, M. Vaez-Iravani,
260. M. Dieterle, G. Mestl, Phys. Chem. Chem. Phys., 4, 822 Appl. Phys. Lett., 60, 2957 (1992).
(2002).
287. G.A. Valaskovic, M. Holton, G.H. Morrison, Appl. Opt.,
261. K. Routray, L.E. Briand, I.E. Wachs, J. Catal., 256, 145 34, 1215 (1995).
(2008).
288. S. Webster, D.A. Smith, D.N. Batchelder, Vibr. Spec-
262. L.E. Briand, A.M. Hirt, I.E. Wachs, J. Catal., 202, 268 trosc., 18, 51 (1998).
(2001).
289. V. Deckert, D. Zeisel, R. Zenobi, T. Vo-Dinh, Anal.
263. I.E. Wachs, Y. Cai, U.S. Patent, US 6245708, 2001. Chem., 70, 2646 (1998).
264. J. Leyrer, D. Mey, H. Knözinger, J. Catal., 124, 349 290. S.R. Emory, S. Nie, J. Phys. Chem. B, 102, 493 (1998).
(1990).
291. B. Humbert, J. Grausem, D. Courjon, J. Phys. Chem. B,
265. P.J. Treado, M.P. Nelson, Handbook of Raman
108, 15714 (2004).
Spectroscopy, eds. I.R. Lewis, H.G.M. Edwards, Marcel
Dekker Inc., New York, Chapter 5, 2001. 292. B. Pettinger, G. Picardi, R. Schuster, G. Ertl, Single Mol.,
3, 285 (2002).
266. L.T.Y. Au, K.L. Yeung, J. Membr. Sci., 194, 33 (2001).
293. B. Pettinger, B. Ren, G. Picardi, R. Schuster, G. Ertl,
267. Y.S.S. Wan, J.L.H. Chau, K.L. Yeung, A. Gavriilidis,
Phys. Rev. Lett., 92, 096101–096101 (2004).
Microporous Mesoporous Mater., 42, 157 (2001).
268. E.S.M. Lai L.T.Y. Au, K.L. Yeung, Microporous and 294. E. Vogel, W. Kiefer, V. Deckert, D. Zeisel, J. Raman
Mesoporous Mater., 54, 63 (2002). Spectrosc., 29, 693 (1998).

269. J. Bergwerff, L.G.A. van de Water, T. Visser, 295. D. Zeisel, V. Deckert, R. Zenobi, T. Vo-Dinh, Chem.
P. de Peinder, B.R.G. Liliveld, K.P. de Jong, B.M. Phys. Lett., 283, 381 (1998).
Weckhuysen, Chem. Eur. J., 11, 4591 (2005). 296. TERS special issue, J. Raman Spectrosc., 38, (2007).
270. K.L. Yeung, N. Yao, J. Nanosci. Nanotech., 4, 1 (2004). 297. A. Urakawa, A. Baiker, Top. Catal., 52(10), 1312–1322
271. K.L. Yeung, X.F. Zhang, W.N. Lau, R. Martin-Aranda, (2009).
Catal. Today, 110, 26 (2005). 298. E. De Smit, I Creemer, J.F. Swart, C. Karunakaran,
272. E. Cao S. Firth P.F. McMillan, F. Gavriilidis, Catal. D. Bertwistle, H.W. Zandbergen, F.M.F. De Groot,
Today 126, 119 (2007). B.M. Weckhuysen, Angew. Chem. Int. Ed. 48, 3632
273. E. Betzig, J. Trautman, Science, 257, 189 (1992). (2009).

274. E. Betzig, J.K. Trautmann, T.D. Harris, J.S. Weiner, 299. A. Urakawa, N. Maeda, A. Baiker, Angew. Chem. – Int.
R.L. Kostelak, Science, 251, 1468 (1991). Ed., 47(48), 9256–9259 (2009).

275. E. Betzig, P.L. Finn, J.S. Weiner, Appl. Phys. Lett., 60, 300. C. Bennici, S.M., Vogelarr, B.M., Nijhuis, T.A., B.M.
2484 (1992). Weckhuysen, Angew. Chem. Int. Ed., 46, 5412 (2007).
276. Ch.S. Fokas, Nachr. Chem. Tech. Lab., 47, 648 (1999). 301. A. Brückner, E. Kondratenko, Catal. Today, 113, 16
(2006).
277. C. Fokas, V. Deckert, Appl. Spectrosc., 56, 192 (2002).
302. U. Bentrup, J. Radnik, U. Armbruster, A. Martin,
278. P. Hoffmann, B. Dutoit, R.-P. Salathe, Ultramicroscopy,
J. Leiterer, F. Emmerling, A. Brückner, Top. Catal., 52,
61, 165 (1995).
1350 (2009).
279. M.P. Houlne, C.M. Sjostrom, R.H. Uibel, J.A. Keimeyer,
J.M. Harris, Anal. Chem., 74, 4311 (2002).
280. C.L. Jahnke, M.A. Paesler, H.D. Hallen, Appl. Phys.
Lett., 67, 2483 (1995).
FURTHER READING
281. C.L. Jahnke, H.D. Hallen, M.A. Paesler, J. Raman D.M. Adams, S.J. Payne, K. Martin, Appl. Spectrosc., 27, 377
Spectrosc., 27, 579 (1996). (1973).
282. S. Münster, S. Werner, S. Mihalcea, C. Scholz, M.A. Bañares, M.V. Martı́nez-Huerta, X. Gao, I.E. Wachs,
E. Oesterschulze, J. Microsc., 17, 186 (1997). J.L.G. Fierro, Stud. Surf. Sci. Catal., 130, 3125 (2000a).
283. W. Noell, M. Abraham, K. Mayr, A. Ruf, Appl. Phys. M.A. Bañares, M.V. Martínez-Huerta, X. Gao, J.L.G. Fierro,
Lett., 70, 1236 (1997). I.E. Wachs, Catal. Today, 61, 295 (2000b).

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034
30 RAMAN SPECTROSCOPY

A. Blume, Ph.D. Thesis, TU, Berlin, 2001. G.J. Puppels, F.F.M. de Mul, C. Otto, J. Greve, M. Robert-
M. Bowden, D.J. Gardiner, G. Rice, D.L. Gerrard, J. Raman Nicould, D.J. Arndt-Jovin, T. Jovin, Nature, 347, 301 (1990).
Spectrosc., 21, 37 (1990). S. Schlücker, M.D. Schäberle, S.W. Huffman, I.W. Levin, Anal.
T.E. Bridges, M.P. Houlne, J.M. Harris, Anal. Chem., 76, 576 Chem., 75, 4312 (2003).
(2004). S.K. Sharma, H.-K. Mao, P.M. Bell, J. Xu, J. Raman Spectrosc.,
P. Dhamelincourt, M. Delhaye, E. Da Silva, J. Raman Spec- 16, 350 (1985).
trosc., 25, 3 (1994). A.V. Talyzin, L.S. Dubrovinsky, T. Le Bihan, U. Jansson, Phys.
D.J. Gardiner, C.J. Littleton, M. Bowden, Appl. Spectrosc., 42, Rev. B, 65, 245413 (2002).
15 (1988). P.J. Treado, A. Govil, M.D. Morris, K.D. Sternitzke, R.L.
P. Gillet, R.J. Hemley, P.F. McMillan, Rev. Mineral., 37, 525 McCreery, Appl. Spectrosc., 44, 1270 (1990).
(1998). P.J. Treado, M.D. Morris, Appl. Spectrosc., 44, 1 (1990).
P.A. Heimann, R. Urstadt, Appl. Opt., 29, 495 (1990). F. Tuinstra, J.L. Koenig, J. Chem. Phys., 33, 1126 (1970).
K.-L.K. Liu, L.-H. Cheng, R.-S. Sheng, M.D. Morris, Appl. D.K. Viers, J.W. Ager, E.T. Loucks, G.M. Rosenblatt, Appl.
Spectrosc., 45, 1 (1991). Opt., 29, 4969 (1990).
S. Merkel, A.F. Goncharov, H.-K. Mao, P. Gillet, R.J. Hemley, I.E. Wachs, J.-M. Jehng, G. Deo, B.M. Weckhuysen, V.V.
Science, 288, 1626 (2000). Guliants, J.B. Benziger, S. Sundaresan, J. Catal., 170, 75
G. Mul, M.A. Bañares, B. van der Linden, S.J. Khatib, J.A. (1997).
Moulijn, Phys. Chem. Chem. Phys., 5, 4378 (2003). M. Watanabe, T. Ogawa, Jpn. J. Appl. Phys, 27, 1066 (1988).
J.C. Petzodt, H. Böhnke, J.W. Gaube, H. Hibst, Proceedings of B.A. Weinstein, G.J. Piermarini, Phys. Rev. B, 12, 1172 (1975).
the 4th World Congress on Oxidation Catalysis, Potsdam, G.H. Wolf, A.V.G. Chizmeshaya, J. Diefenbacher, M.J.
Germany, Supl. 16, 2001. McKelvy, Env. Sci. Technol., 38, 932 (2004).
G.J. Puppels, W. Colier, J.H.F. Olmikhof, C. Otto, F.F.M. de
Mul, J. Greve, J. Raman Spectrosc., 22, 217 (1991).

Encyclopedia of Analytical Chemistry, Online © 2006–2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Analytical Chemistry in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470027318.a9034

You might also like