Sunteți pe pagina 1din 52

4

Applications of Raman Spectroscopy to


Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

Heterogeneous Catalysis
BY ISRAEL E. WACHS AND FRANKLIN D. HARDCASTLE

1 Introduction
Downloaded by McGill University on 26 December 2012

During the past few years Raman spectroscopy has become a powerful characteriza-
tion technique in the field of heterogeneous catalysis. The dramatic increase in the
research activity of Raman spectroscopy in heterogeneous catalysis is due to the unique
molecular level structuralinformation provided by Raman experiments as well as the rapid
advances in Raman instrumentation. State-of-the-art Raman spectrometers offering
superior stray-lightrejection capabilities and optical multichannel detectors, with quantum
efficiencies ranging from 10% to 30%, have been developed. These developments in
instrumentation, coupled with increased efficiency in hardware and software for data
manipulation, have significantly altered the range of problems that can be investigated
with Raman spectroscopy.
In many instances Raman spectroscopy can address important catalytic problems
that cannot be examined by other spectroscopies. For example, the extremely low Raman
scattering from water allows for the direct monitoring of molecular events that occur in
aqueous solution during catalyst synthesis. The weak Raman signals of support materials
such as alumina and silica result in a weak substrate background spectrum and allow for
the observation of the vibrational spectra of absorbed molecules and supported metal
oxides below 1200 cm-'. In fact, Raman spectroscopy is uniquely suited for the
characterization of supported metal-oxide catalysts because it is able to discriminate
among the various metal-oxide phases simultaneously present in such multicomponent
systems as well as provide molecular level structural information. Another advantage of
the Raman technique is that experiments can be performed in situ with no stringent
requirements on the sample preparation nor on the sample environment within the cell.
The recent advances in Raman instrumentation have also led to the emergence of Raman
spectroscopy as a surface-science analytical technique (i.e., the detection and quantification
of very low concentrationsof molecules on low surface area single crystal surfaces). The
full range of problems amenable to modem Raman spectroscopic analysis that face the
catalytic scientist is just beginning to be realized.
A number of reviews have recently appeared in the literature concerning the Raman
spectroscopy of catalytic systems.14 The purpose of this chapter is not to review the
literature, but to demonstrate what is currently achievable in Raman spectroscopy of
heterogeneous catalysis. The theory of the Raman effect and its relationship to signal
intensity are presented in Section 2. Point-group symmetry and the Frequency/Length/
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 103

Strength (FLS) empirical approach are discussed in Section 3 as methods for determining
the molecular structures of metal-oxide species from their Raman spectra. The state-of-
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

the-art in Raman instrumentation as well as new instrumental developments are discussed


in Section 4. Sampling techniques typically employed in Raman spectroscopy experi-
ments, ambient as well as in situ, are reviewed in Section 5. The application of Raman
spectroscopy to problems in heterogeneous catalysis (bulk mixed-oxide catalysts,
supported metal-oxide catalysts, zeolites, and chemisorption studies) is discussed in depth
in Section 6 by selecting a few recent examples from the literature. The future potential
of Raman spectroscopy in heterogeneous catalysis is discussed in the final section.
Downloaded by McGill University on 26 December 2012

2 Theorv of the Raman Effect

Observations of the Raman effect were first reported in 1928 by Raman and
Krishnan in the course of work carried out on the light-scattering ability of molecules5:
If we assume that the x-ray scattering of the unmodified observed by
Prof. Compton corresponds to the normal or average StatetOyFethe atoms and
molecules, while the modified scattering of altered wave-length corresponds
to their fluctuations from that state, it would follow that we should ex t
also in the case of ordinary light two types of scatterin one determinsy
the normal o tical properties of the atoms or mokules, and another
representing tKe effect of their fluctuations from their normal state. It
accordingly becomes necessary to test whether this is actually the case. The
experiments we have made have confiied this antici ation, and shown that
in every case in which li ht is scattered b the molecufes in dust-free liquids
or gases, the diffuse raJation of the or&ary kind, having the same wave
length as the incident beam, is accompanied by a modified scattered
radiation of degraded frequency.
C.V. Raman and K.S.Krishnan
Feb. 16, 1928
Scattered light can be thought of as consisting of an elastic component (Rayleigh
component, or the "diffuse radiation of the ordinary kind"), and an inelastically scattered
component referred to by Raman as "a modified scattered radiation of degraded
frequency." This modified radiation is commonly referred to as Raman scattering. Raman
scattering may be regarded as an inelastic collision of an incident photon with a molecule
that is either in its ground electronic state, giving rise to Stokes Raman lines (degraded
frequency), or in an electronically excited state which gives rise to anti-Stokes Raman
lines (enhanced frequency). In contrast with fluorescence and phosphorescence, the
Raman effect does not require the incident light to be coincident with an absorption
transition because the photon is never entirely absorbed. Instead, the photon merely
perturbs the molecule, thereby inducing a vibrational or rotational transition. In principle,
this perturbation may be induced by any photon energy and, consequently, any wavelength
of light may be used to obtain a Raman spectrum.
The Raman effect depends on the polarkability of the molecule and the dipole
moment induced or distorted by the electric field of the incident light.6 Given the
View Online
104 Catalysis

relatively large masses of the atomic nuclei and incident light of high enough frequency,
the exchange of energy between the incident photon and the rotational or vibrational
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

motions of the molecule results almost entirely from the polarizability of the electrons.
The selection rule for a vibrational Raman transition depends on whether the polarizability
changes when the molecule vibrates in a certain way. If the polarizability changes with
respect to nuclei displacement, then that particular vibration is Raman active. For a linear
polyatomic molecule composed of N nuclei, the molecule may vibrate 3N-5 ways,
whereas a nonlinear polyatomic molecule may vibrate 3N-6ways. Consequently, the
linear and nonlinear polyatomic species are said to possess a maximum of 3N-5and 3N-6
Downloaded by McGill University on 26 December 2012

fundamental vibrational modes, respectively.


The induced electric dipole moment P responsible for the Raman signal can be
related to the electric field E of the incident radiation by the power series
P = P(')+P(2)+...
where P(') = aE,P(2)= BEE/2, and so on, where a and B are constants? The first-order
term P(') possesses the correct frequency dependence for normal Raman scattering while
the second and subsequent terms give rise to nonlinear Raman effects (e.g., hyper-Raman
scattering). In general, especially if the incident flux of the laser beam is not too
excessive, truncating the series after the linear term is justified because this series rapidly
converges. This leads to the following expression relating the induced dipole moment P
to the electric field E for normal Raman scattering:*
P=aE
where a is defined as the molecular polarizability tensor and P and E are vectors. The
polarizability of the molecule is a measure of how readily the electrons are perturbed by
the incident field. Thus, the magnitude of the induced dipole moment depends on the
amplitude of the incident laser field and the polarizability of the molecule. In general, for
the anisotropic case, the components of the induced dipole moment can be expressed by
the following set of equations:
p x = %Ex + %yEy + Q2
p y = %,Ex + aypy + 4 . 8 2
p2 = %,Ex + %yEy + %A-
The polarizability tensor is defined by an array of nine components but, in most cases, the
tensor is symmetric so that %y = %, ayz= %, and = azX.The resulting six
components of the polarizability tensor can be better visualized by considering an equation
for an ellipsoid in Cartesian coordinates:
%X2 + 4,,y2 + %Z2 + 2%yxy + 25,yz + 2Clyzyz + 2 s z x = 1.
The ellipsoid described by the above equation conforms to the shape of the electron clouds
surrounding the entire molecule. The ellipsoid may be further regarded as representing
the polarizability because vibrational modes which induce a change in the size or shape
of the polarizability ellipsoid are considered to be Raman active modes.
The classical theory of Raman scattering offered a mechanistic way of understand-
ing the Raman effect but was unable to provide information on the intensity of Raman
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 105

transitions. A polarizability theory based on quantum theory and describing the scattering
intensity for the normal Raman effect was developed by Placzek in 1934.9 The
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

polarizability theory predicted experimentally determined intensities of both the Stokes and
anti-Stokes intensities. There are a few assumptions incorporated into this theory,
however, which limit its applicability to the ordinary or normal Raman effect and not, for
example, the resonance Raman effect. First of all, the excitation frequency must be larger
than the frequency associated with any vibration or rotation transition. Next, the
excitation frequency must be much less than any electronic transition (to avoid electronic
resonance or preresonance effects), and finally, the ground electronic state must not
Downloaded by McGill University on 26 December 2012

normally be degenerate. The polarizability theory predicts that the Raman intensity, I,
depends not only on the square of the incident laser field, E, but also on the square of the
derivative of the polarizability with respect to a normal vibration coordinate, q, or
I = (Sa/6q)2 E2.
Another useful result is the predicted relationship between the relative intensities of the
Stokes and the anti-Stokes Raman lines given by
'anti-StokesfiStokes = e-hV/kT (vo + Vl4 /(vo - Vl4
where v, is the frequency of the Rayleigh line (incident energy) and v is the frequency
location of the Raman line in wavenumbers. This relationship between the intensities of
the Stokes and the anti-Stokes Raman lines can serve as a rather expensive thermometer
because the experimental variable is the temperature, T, which determines the magnitude
of the intensity ratio. This relationship also describes the "fourth-power" dependence of
the Raman intensity with respect to the frequency region utilized. Stated simply, the
"fourth-power"or v4 rule says that the Raman signal should increase as the frequency of
the incident laser beam is increased. Based on this assumption, and in the complete
absence of other effects, an ultraviolet or visible laser should give a significant
enhancement in Raman intensity over an infrared laser source.

3 The Raman Effect and Molecular Svmmetrv

3.1 Point-Grow Svmmetrv. - Plazcek's polarizability theory explains that the intensities
of the Raman bands depend on the change in the polarizability of the electronic wave
functions as the molecule vibrates. However, what are the selection rules that determine
the number of allowed Raman bands? It turns out that the selection rules for vibrational
modes allowed for a particular molecule are based entirely on the spatial orientation or
molecular geometry of that molecule. The molecular geometry can be characterized by
a set of symmetry elements that describe the geometry of the molecule and the allowed
translational, rotational, and vibrational (both Raman and infrared) modes for the
molecule. The symmetry operations (rotations, reflections, etc.) that define the geometry
of the molecule belong to the so-called "point group" of symmetry operations characteris-
tic of that particular geometry.
View Online
106 Catalysis

The vibrational information indicating a specific molecular geometry is tabulated


in "character tables" which are readily available in the appendices of many excellent
'
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

accounts on the point-group theoretical treatment of simple molecules.6910*'The character


table is essentially a transformation matrix that is classified by the overall symmetry of
the molecule (point group) and contains the symmetry elements of the point group along
the first row and the types of representations or modes along the first column. The type
of representation can describe a nondegenerate symmetric vibration, denoted as A, or a
nondegenerate antisymmetric vibration, denoted as B. If the vibration is degenerate, then
other symbols are used such as E for a doubly degenerate vibration and F (or T) for a
Downloaded by McGill University on 26 December 2012

triply degenerate vibration. Most character tables are "user-friendly" and quickly identify
modes as Raman active once these modes are determined. The determination of these
modes, or symmetry species, involves calculating the symmetry species spanned by the
molecule of interest by applying the "Great Orthogonality Theorem" (GOT) or, more
appropriately, the "Little Orthogonality Theorem" (LOT) which is a ready-to-use formula
handed down from the group theorists. The character table can then be used to determine
which of the modes are vibrations (vs. rotations and translations), as well as which are
Raman active and/or infrared active. Thus, if the structure for a given molecule were
known, a thorough application of vibrational mode analysis could be used to determine
the number of permitted vibrational frequencies as well as the entire Raman spectrum."
Again, this procedure is exhaustively covered in most relevant texts on the subject (see,
for example, references 6, 10, and 11) and thus will not be presented here.
The discrimination between alternate models of differing molecular geometries that
might be proposed for a given molecular species becomes a viable task when vibrational
modes are assessed. Consider an undistorted tetrahedral metal oxide species Mo4 (Td
symmetry group; 0 = oxygen and M = metal cation). A group-theoretical analysis on the
sigma and pi orbitals of the MO, tetrahedron gives the irreducible representation of the
tetrahedral group as r = A, + E + F, + 3F2.10*11The character table for the Td point
group shows the translations to be attributed to the triply degenerate F2 symmetry species
and the rotations to the F, symmetry species. This leaves four fundamental vibrational
modes, rvib = vl(Al) + v2(E) + ~3(F2)+ vq(F2) and is consistent with the total possible
expected vibrations: 3N-6 = 3(5)-6 = 9 modes, where N is the total number of atoms in
the molecular species (recall that the A notation represents single modes, whereas E
indicates doubly degenerate modes and F (or T) triply degenerate modes). All of these
fundamental vibrational modes are Raman active as illustrated by the change in
polarizability for the particular symmetry species (typically shown in the last column of
the character table). As the symmetry of the molecule is lowered (for example, C 3 , C2",
and so on), either by chemically substituting groups or by exposure to external fields,
mode-splitting occurs and more of the nine fundamental modes (3N-6) become Raman
active due to the removal of degeneracies.
For an undistorted octahedral metal-oxide molecule MO, (Oh point-group
symmetry), a treatment similar to that for the M04 tetrahedron may be carried out using
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 107

the 0, character table. This analysis yields three Raman active vibrational modes for the
perfect M06 octahedron: r,, = vI(AIg)+ v2(Eg)+ ~3(F2~). As distortions occur and the
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

high degree of symmetry is lost, a progressively greater number of the 15 possible


fundamental modes (3N-6 = 3(7)-6 = 15) become Raman active. As a final note, the
Raman spectrum of a metal-oxide system may be further complicated by other effects
(combination bands, overtones, or lattice modes) that are not regarded as fundamental
mode splitting, and this may result in a greater number of observed bands than would
normally be expected.
The application of simple point-group theory has provided detailed molecular
Downloaded by McGill University on 26 December 2012

information about supported metal-oxide phases' and will be discussed later in the section
on applications of Raman spectroscopy (Section 6). The determination of the structure
of transition metal-oxide species in heterogeneous catalysts, however, requires more than
a group theoretical approach. A comprehensive comparison of the Raman spectrum of the
unknown species with those of reference compounds having known structures must be
performed. It is particularly appropriate to utilize reference compounds when studying
such nonideal compounds as crystalline metal-oxide lattices (where the metal-oxide species
is distorted by the three-dimensional lattice) or two-dimensional surface-oxide species
(which are usually distorted by their interaction with the oxide substrate). A comparison
of the Raman spectra of species possessing an unknown structure with the spectra of
species possessing known structures leads to the so-called "fingerprint approach" of
determining the molecular structures of metal oxides.

3.2 The FLS Amroach: Raman Stretching FreuuencvBond LengthBond Strength


EmDirical Correlations. - The choice of method used to interpret the Raman spectrum may
often be the major obstacle in determining the molecular structure of a metal oxide
species. The most widely accepted method of ascertaining molecular structures from
Raman spectra, as mentioned in the previous subsection, is the so-called "fingerprint"
approach. As its name implies, the fingerprint approach involves a direct comparison of
the Raman spectrum for the unknown molecule with those of one or more reference
compounds. In most cases, however, the Raman band positions and relative intensities
of the unknown structure do not precisely match those of reference structures. The
investigator must then evaluate the significance of this discrepancy and, consequently,
subjectivity often plays a major role in the interpretation of the Raman spectra of metal
oxides; this type of subjectivity often leads to erroneous or contradictory conclusions.
Nevertheless, the fingerprint approach has proven quite successful for geometrically
symmetric or very regular metal-oxide structures. For example, the hydrated two-
dimensional surface rhenium oxide supported on yA1203 (180 m2/g) was shown to consist
of isolated, slightly distorted ReO, tetrahedra (q, point-group symmetry),12and surface
chromates on yAl2O3,TiO, (Degussa, P-25; 60 m2/g), and SiO, (Cab-0-Sil, 325 m2/g)
supports were found to consist of tetrahedral monomers, dimers, and trimers.13 The
coordination number and local symmetry were found by comparing the Raman spectra of
View Online
108 Catalysis

the hydrated surface species with those of structurally similar species in aqueous solution.
Once the moisture was removed from these systems, however, the surface rhenates and
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

chromates became very distorted as a result of increased interaction with the support as
indicated by a dramatic upward shift in frequency of the metal-oxygen stretching mode.
Such distortions from ideal symmetry are also particularly prevalent for metal-oxide
species confined to the solid state, where distortions are imposed by either the crystalline
lattice (long-range order) or by immediate surroundings (short or medium-range order).
Hence, the fingerprint approach is of limited usefulness because of the limited number of
available reference compounds and the virtually infinite number of possible structural
Downloaded by McGill University on 26 December 2012

distortions available to the unknown structure, These shortcomings have generated the
need for a more systematic and reliable method for determining the molecular structures
of transition metal oxides from their Raman spectra.
Recently, a method was developed for interpreting Raman spectra that allows the
systematic determination of bond lengths and coordination numbers for several types of
transition metal oxides. This method, labeled as the "FLS Approach," relies on empirical
relations between metal-oxygen Raman stretchingfrequencies,corresponding metal-oxygen
bond lengths, and Pauling bond strengths. The observed Raman stretching modes of a
metal-oxide species (in most cases > 400 cm-') are converted to bond strengths, in valence
units, and bond lengths in angstroms. The bond strengths are added in a systematic
fashion, and all combinations that add to the formal oxidation state of the metal cation,
according to the valence sum rule, represent viable molecular structures; typically, one of
these structures prevails as the most reasonable structure. Empirical Raman stretching
frequencyhnd length relations have been established for several metal-oxide systems
including molybdate~,'~ van ad ate^,'^ niobates,16 bismuth oxides,17 tungstates,'* and
titanates.'' In each case, the general form of the relation was found to be
v = A exp (- B R)
where A and B are fitting parameters, v is the metal-oxygen Raman stretching frequency
in wavenumbers, and R is the metal-oxygen bond length in angstroms. In addition to
empirical frequencyhength relations, Pauling bond strengthhength relations are also used
in the FLS approach. Pauling bond strengths, also referred to as bond orders or bond
valences, are indispensable tools for discussing the electronic structure of proposed metal-
oxide species.20 The bond sbength reflects the relative distribution of available valence
electrons throughout the covalent bonding of metal-oxide species. According to the
valence sum rule?' the total valence of a metal cation is conserved. Hence, the sum of
the individual metal-oxygen bond strengths of a metal-oxide species, Esi, may be equated
to the formal oxidation state of the metal cation, V, according to
zsi = v.
A generally applicable relationship has been developed by Brown and Wu for
relating Pauling bond strengths to bond lengths for metal-oxygen bonds:
s = (R/C)-D,
where C and D are fitting parameters, s is the Pauling bond strength in valence units (vu),
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 109

and R is the bond length in angstroms?2 The empirical parameters for the Raman
stretching frequency/bond length relation (A and B) and the bond strength/bond length
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

relation (C and D) are tabulated in Table 1 for bismuth-, titanium-, vanadium-, niobium-,
molybdenum-, and tungsten-oxygen bonds. The FLS approach utilizes these empirical
correlations and provides a systematic method for determining the coordination numbers
and bond lengths for these metal oxides.

Table 1 Empirical parameters for Raman stretching frequency/bond length and


bond strengthhength correlations
Downloaded by McGill University on 26 December 2012

v = A exp(-BR) s = (WC)'D
M - 0 bond A B C D
~~~~~

Bi-0 92,760 2.511 2.010 5.0


Ti-0 13,322 1.617 1.806 5.2
v-0 21,349 1.918 1.791 5.1
Nb-0 25,922 1.917 1.907 5.0
Mo-0 32,895 2.073 1.882 6.0
w-0 25,823 1.902 1.904 6.0

One interesting application of the FLS relations is the ability to easily estimate the
Raman stretching frequencies of perfectly symmetric metal-oxide species. For example,
the Raman stretching frequencies for the perfect MOO, tetrahedron and MOO, octahedron
may be estimated from the FLS relations and empirical parameters from Table 1. First,
the 6.0 valence units (vu) of available charge of each Mob metal cation is equally divided
among the total number of Mo-0 bonds; this is because each Mo-0 bond per molybdate
species is identical due to its high degree of symmetry. This results in a valence of 6/4
= 1.5 vu per Mo-0 bond for the MOO, tetrahedron and 6/6 = 1.0 vu per bond for the
MOO,octahedron. These bond strengths may be converted to bond lengths of 1.759 and
1.882 A for the tetrahedron and octahedron, respectively, by using the appropriate
equation and parameters from Table 1. The Mo-0 Raman stretching frequencies for these
bonds may be obtained by combining the frequencybength and strengthbength relations;
this results in an Mo-0 stretch of 858 cm-' for the perfect MOO, and 655 cm-' for the
perfect Moo6.
The Raman stretching frequencies for theperfect MOO, and MOO, structures may
be compared with those of molybdate reference compounds. The most suitable reference
MOO, tetrahedron is identified in PbMoO, which has its highest occurring stretching
frequency at 869 cm-l,14only 11 cm" higher than that estimated for the perfect structure
at 858 cm-' (about one standard deviation of error associated with the FLS parameters).
The most regular reference MOO, structure is represented by the slightly distorted
octahedron present in the perovskite Ba+MoO,, which has its highest occurring Mo-0
View Online
110 Catalysis

stretching frequency at 813 cm-1,23*24or 147 cm-' higher than that estimated for the
perfect MOO, at 665 cm-', reflecting a significant distortion present in this structure. The
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

estimated Raman stretching frequencies of perfect structures for several metal-oxide


systems are tabulated in Table 2. Applications of the FLS approach for systematically
determining the coordination numbers and bond lengths of molybdates are presented in
Section 6 for bismuth molybdate catalysts.
~~

Table 2 Estimated metal-oxygen Raman stretching frequencies (cm-') for perfect


tetrahedral and octahedral metal-oxide structures
Downloaded by McGill University on 26 December 2012

Metal cation MO4 M06

Bi3' 442 28 1
Bi5' 743 493
Ti4' 724 572
v5+ 796 607
Nb5+ 787 586
Mo6' 858 665
w6' 874 69 1

4 Raman Instrumentation

4.1 Lasers. - The instrumentation comprising the state of the art in Raman spectroscopy
has progressed at a remarkable rate in recent years. Modern day Raman spectroscopy has
been commonly referred to as "laser" Raman spectroscopy to differentiate from the older
method of inducing the Raman effect with mercury-arc lamps (emission lines at 435.8 and
546.1 nm). Lasers offer a notable advantage over arc lamps as excitation sources because
the emission from a laser offers a highly intense beam of light that is spatially coherent
and highly monochromatic. The lasers most commonly used in Raman spectroscopy are
those that emit in the visible region of the spectrum. Gas lasers such as the helium neon
(632.8 nm), argon ion (514.5 and 488.0 nm), and krypton ion (752.5, 647.1, 530.9, and
568.2 nm) are among the most popular gas lasers used today. Solid-state lasers such as
the ruby (@-doped, a-alumina), the Nd:YAG (neodymium-doped yttrium aluminum
garnet), and semiconductor lasers have also been used in Raman studies as excitation
sources in the visible region. Each of these laser systems emits a specified set of
wavelengths defined by the allowed transitions of the active medium in the resonator
cavity of the laser.
If a greater degree of tunability in excitation wavelength is required, a tunable dye
laser may be used. The tunable dye laser is typically pumped with an intense beam
originating from another laser (e.g.; Nd:YAG or argon ion) or from a xenon-filled
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 111

flashlamp. The solvated dye molecules are irradiated with ultraviolet or visible radiation
which populates the ro-vibronic levels of the first excited singlet state. The excited singlet
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

state can then undergo rapid radiationless transitions to a triplet state via intersystem
crossing or may radiate back to the ground state. The populated state "fluoresces" back
to the ground state producing a red-shifted fluorescence spectrum. The dye itself
possesses many fluorescing transitions and emits an intense and broad-band fluorescence
spectrum over a range of frequencies in the visible spectrum. The frequency range of this
emission depends on the type of dye being used. Perhaps the most popular laser dye is
Rhodamine 6G which fluoresces strongly from the yellow to the red region of the visible
Downloaded by McGill University on 26 December 2012

spectrum: -560 to 660 nm. In practice, wavelength selection is generally accomplished


with a prism, but gratings or filters can also be employed. In general, the tunable dye
laser permits essentially unrestricted and continuous tunability of the Raman excitation
source over the entire visible region. This flexibility in the excitation wavelength is
desirable when Raman intensities are to be correlated with the excitation wavelength, or
when resonant Raman enhancement of a particular species is sought.

4.2 Monochromators. - After collecting the scattered light emitted from the sample, the
problem of separating the Raman component from the Rayleigh and quasi-elastically
scattered components remains. Raman scattering is an extremely inefficient process
compared to that of light absorption, and the intensity of a Raman band is sometimes on
the order of times that of the excitation source. Thus, the stray light rejection
capability of the dispersing system (i.e., prisms, gratings, etc.) must be very efficient,
especially when the vibrational spectra of catalytically active surface metal-oxide species
are sought. The dispersing elements found in a modem Raman spectrometer are
diffraction gratings. Although single grating monochromators are available, Raman
spectrometers usually require at least two gratings (unless a stray light rejection filter is
used). Double and triple monochromators are commercially available to obtain a
sufficient d e p of rejection of the stray light.
A major concern in any Raman experiment is the presence of "ghost" lines
(extraneous peaks) appearing in the Raman spectrum when a high Rayleigh background
is present. The ghost lines occur at the same wavenumber position from spectrum to
spectrum (given identical experimental settings) and result from the presence of
imperfections in the construction of the diffraction gratings. Ghost lines may be subdued
by increasing the number of gratings in the system; i.e., by using a double monochromator
or even a triple monochromator where ghost lines are virtually eliminated. In a typical
single monochromator about lo-' of the radiation may appear as a ghost line. For a
double monochromator only (10-5)2= lo-'' of the radiation may be present as the ghost
line,and for a triple monochromator this fraction is reduced to = lO-".'
A double monochromator has two separate gratings which are separated by an
adjustable slit. Both of these gratings are mechanically coupled onto the same shaft so
that tracking errors are virtually eliminated. The gratings are rotated concurrently and the
View Online
112 Catalysis

instrument scans as each resolution element (band pass) of the Raman spectrum is
successively detected photoelectrically as a function of time. One double monochromator
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

used successfully to collect Raman spectra has 1800-grooves/mm gratings which are
holographic and not ruled, thereby decreasing the possibility of observing ghost peaks due
to grating imperfections (SPEXIndustries, Inc., Metuchen, NJ). The gratings are coupled
and mounted in a Czerny-Turner additive dispersion mounting (two separate spherical
mirrors are placed side by side and opposite to the grating, with the slits located on either
side of the grating). The focal length of the spectrometer is 0.85 m and the aperture
possesses an f-number of 7.8 (f/7.8). The resolution is 0.15 cm-' and the reproducibility
Downloaded by McGill University on 26 December 2012

of the measurement is 0.2 cm-'. Because this is a scanning spectrometer, the time scale
is calibrated to a wavenumber scale or the wavenumber is read directly from a scale which
relates to the position of the gratings to the wavenumber scale.
The triple-grating spectrometer has become the state of the art in Raman
spectrometers over the past several years because of its excellent stray-light rejection
capabilities and flat, undistorted focal plane projected at the exit slit. These valuable
characteristics enable the routine collection of very weak signals which have previously
been overwhelmed by the intense Rayleigh background. The flat, undistorted focal plane
projected at the exit slit allows the triple monochromator to be used in the spectrograph
mode (i.e., a wide range of wavenumbers reach the exit slit simultaneously as, for
example, when a photographic plate is the detector). A triple monochromator is
essentially a two-stage spectrometer (spectral coverage 185 to lo00 nm, accuracy 0.5 nm,
resettability 0.2 nm, stray light at 10 bandpass units from the laser line). The fxst
stage consists of 0.22-m double monochromator with coupled concave gratings forming
a subtractive-dispersion stage. The second stage is a 0.6-m single-monochromator
spectrograph with a turret which provides the convenience of using one of three gratings
possessing different groove densities (typically 600, 1200, and 1800 grooves/mm). The
grating selection depends on the desired resolution and spectral range. The first stage of
the spectrometer acts as a variable wavelength, selectable bandpass filter which removes
the Rayleigh scattering and much of the unwanted or stray light from the signal, and
refocuses the resulting inelastic scattering through the entrance slit of the spectrograph
stage. The spectrograph stage can either disperse the entire desired spectral region of the
Raman spectrum onto the focal plane of an intensified photodiode array (IPDA) or charge-
coupled device (CCD) detector/optical-multichannel-analyzer(OMA) where all resolution
elements of the spectrum are collected simultaneously, or it can operate in the scanning
mode and disperse one resolution element of the spectrum at a time onto a photoelectric
detector.

4.3 Detectors. - Once the Raman signal has been successfully extracted from all other
components of the scattered radiation, the Raman signal impinges upon a detector. The
type of detector chosen depends on the optical arrangement of the monochromator and the
demands of the spectroscopist. If the monochromator is to be used as a spectrograph, then
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 113

the detector must mimic a photographic plate where an image of the entire spectrum is
projected onto a photodiode array (OMA), and all resolution elements of the spectrum are
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

collected simultaneously. When a monochromator is used as a spectrometer, then the


detector is single channel and one resolution element (or small section) of the Raman
spectrum is detected at any one time as the grating is rotated or scanned as a function of
time.
Modem detectors for Raman spectrometers almost invariably function as photon
counters because this method of detection is far more sensitive for differentiating
extremely low-level light signals from the random noise inherent in the mechanism of the
Downloaded by McGill University on 26 December 2012

detector. Photon counting is used in photomultipliers and in intensified linear photodiode


arrays. When a photon collides with a photosensitive cathode of a photomultiplier, the
cathode emits a photoelectron which is accelerated to the first dynode in the photomulti-
plier tube through a substantial voltage difference. The fust dynode responds to this
collision by emitting several new electrons, and these are accelerated to the next dynode
where more electrons are produced. This "cascading" effect from successive dynode
stages results in an easily detectable pulse consisting of about lo6 electrons for a typical
photomultiplier tube (12 stage). The resulting current produced by this pulse is
subsequently a.c.-amplified for digital (or analogue) output. Thus, in principle, each
photon signal is amplified and counted as a giant pulse of electricity.
The popularity of the triple monochromator in Raman spectroscopy is owed almost
entirely to the popularity of solid-state linear IPDAs and CCD detectors. The triple
monochromator/OMA Raman spectrometer is capable of simultaneously recording all the
spectral resolution elements of a Raman spectrum without mechanically rotating the
spectrometer's grating (scanning). Thus, the multiplex advantage is realized by the
capability of obtaining spectra with superior signal-to-noise ratios in only a fraction of the
acquisition time required of the more conventional scanning instruments. Additionally,
the mechanical backlashing errors associated with the movement of the cosecant arm of
the scanning grating are absent in the OMA arrangement because the grating positions are
held constant. Typically, the improvement in acquisition time for the OMA is on the
order of 1000 times with an associated improvement in the signal-to-noise ratio of over
30 (assuming 1000 resolution elements). For a routine sample run in the authors'
laboratory, the typical time scale involved in collecting a Raman spectrum for a surface
metal oxide system with an OMA is on the order of 5 minutes for a 1000-cm-' spectral
region (1.5 cm-I resolution) scan. A Raman spectrum with a similar signal-to-noise ratio
collected on a scanning instrument could take as long as 2 days. The much faster
acquisition time of the OMA also makes time-resolved Raman spectroscopic observations
tenable.
The optical multichannel photodiode array most used in Raman spectroscopy is the
IPDA consisting of a one-dimensional array of amplified photodiodes. The mechanism
behind the JPDA detection is that each photodiode converts photons to separated electron-
hole pairs (semiconductor-amplification device). In some solid-state detectors (e.g.,
View Online
114 Catalysis

EG&G, Princeton, NJ), the photodiode is fabricated with a multialkali photosensitive


material on quartz which is coupled with fiber optics to a proximity-focused microchannel
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

plate intensifier. Some instruments have an 18-mm intensifier possessing either 512 or
1024 (-700 usable) elements in the array, while others have a 25-mm intensifier with
1024 (-lo00 usable) elements in the array. In some of these detectors, quantum
efficiencies of over 10% are achievable, and this makes the sensitivities of the IPDA
detectors comparable to that of a photomultiplier tube. These IPDA detectors are capable
of being gated from 10 ns to 10 ms with an on/off ratio of >1@. Cooling of the IPDA
is accomplished thermoelectrically to -40 "C (aided by circulating refrigerated ethylene
Downloaded by McGill University on 26 December 2012

glycol/water at -25 "C) to minimize the noise contribution due to the diode dark count.
The IPDA detector may be controlled by a detector interface and OMA console which
offers menu-driven software and touchscreen/keyboard operation. For the collection of
extremely low-light-level Raman signals (0.05 photoelectrons per second per pixel)
adjustments must be made to the electronics of the interface board (intensifier gain, clock
timing, and pixel gain response to light) to ensure the most efficient and noise-free
communication between the OMA and the detector. Further cooling (liq.-N2) and
insulating of the detector are sometimes employed to minimize the equivalent background
illumination (EBI) noise from the intensifier. Finally, corrections can be made to the raw
data including background subtraction, pixel-gain correction, spectrometer throughput, and
selection of the optimized integration timeF5
Although the IPDA is presently the most popular detector used in Raman
spectroscopic studies, CCD detectors are rapidly replacing IPDAs for many applications.
CCDs consist of a two-dimensional array of pixels, usually 512 X 512, but with 2 X 1024
being the most popular size for Raman applications. The mechanism for photon detection
by the CCD is different from that of an IPDA?6 In a CCD, the photons are converted to
electron-hole pairs, and the electrons are collected into a potential well created by a
depletion region, and this is read as a net charge. The additive charge contributions for
each column of pixels is determined in a process called binning which increases the
signal-to-noise ratio. One advantage of CCDs over IPDAs is that the dark current is
virtually eliminated by cooling the CCD to liquid nitrogen temperatures, and this results
in a much higher signal-to-noise ratio. Another advantage is that quantum efficiencies of
over 35% may be achieved with CCDs in the red region of the spectrum; red excitation
(or infrared excitation) is one technique of removing the effects of broad-band fluores-
cence from the Raman spectrum.

4.4 Raman Systems. - Raman instruments may be constructed by assembling components


piece by piece, or one may rely on the experience and (business-driven) sense of Raman
manufacturers/distributorsand acquire a Raman "package." Figure 1 shows the piece by
piece Raman spectroscopy apparatus used in the authors' laboratory. The laser source is
a Spectra-Physics argon-ion laser (Model 165), the monochromator is a SPEX Triplemate,
and the detector is an IPDA/OMA (EG&G, PAR) which is controlled with a dedicated
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 115
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

Argon ion laser

filter I!I detector


Y7K T
OMA Ill
Downloaded by McGill University on 26 December 2012

I 1
illuminator

Plotter
Spectrometer

Figure 1 Schematlc of Raman spectroscopy apparatus.

OMA III console control panel (Motorola microprocessor). The data manipulation may
be performed using the OMA integrated software package, or the data may be reformatted
to MS-DOS and imported to popular PC-based (e.g., Lotus 1-2-3) software packages for
further analysis. The OMA to MS-DOS reformatting capability resides in the OMA III
software package.

4.5 Fluorescence and Raman Instrumentation. - In many cases, the biggest problem
facing the Raman spectroscopist may not be the noise or detection limits of the Raman
instrument, but the luminescence/fluorescence background emanating from the sample or
impurities present in the sample. The Raman scattering process is inefficient compared
with that of fluorescence, and often overwhelms the Raman signal. As a general indicator,
for an incident flux of 108 photons making contact with the sample only one photon will
be Raman scattered. Fluorescence, however, which is a resonant phenomenon, will
typically be 10 times as efficient as the Raman effect and yield a cumbersome broad-band
background emission spectrum that is superimposed onto the Stokes lines of the Raman
spectrum. Although the problem of fluorescence deters the collection of Raman data for
many types of samples, there are instrumental approaches used to discriminate against the
fluorescence background.
4.5.1 Time-Resolved Techniaues. Early attempts were made at discriminating the Raman
signal from the fluorescence background by time-resolved techniques. The Raman effect
is a vibrational process and occurs on the time scale of about sec, while fluorescence
lifetimes typically occur in the nanosecond to picosecond regime. Because the time scales
View Online
116 Catalysis

generally differ by a factor of 100, it should in principle be possible to discriminate


between the two effects. A pulsed, frequency-doubled Nd:YAG laser and nanosecond
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

photon-counting detection system have been used to reduce the relative intensity of the
fluorescence background for an Mn-doped ZnSe sample.” Advances in continuous wave
(cw)-laser technology have allowed the mode-locked cw laser to be used in the time-
resolved spectroscopic discrimination against much of the fluorescence background.28 A
mode-locked, cavity-dumped argon-ion laser has also been used to obtain the sub-
nanosecond time-resolved rejection of fluorescence from the Raman spectrum of solutions
of benzene doped with the fluorophors acridine orange and n1brene.2~
Downloaded by McGill University on 26 December 2012

4.5.2 Multidexine Methods. The most straightforward method of avoiding the


fluorescence background, however, is simply to avoid exciting the fluorescing transitions
by changing the excitation wavelength to either the ultraviolet or to the red or near
infrared where most luminescing/fluorescing transitions are avoided. Ultraviolet radiation,
however, can cause serious sample degradation, and the Raman signal in the near infrared
is seriously degraded because of the v4 frequency dependence of the Raman intensity. In
addition to the loss of Raman intensity due to its v4 dependence in the red or near-infrared
region, conventional Raman detectors have notoriously low responses in this spectral
region, thereby compounding the problem. The recent commercial availability of CCD
detectors, however, with very high quantum efficiencies in the red (above 30%),has aided
significantly in Raman excitation using low-frequency light sources?’
Recently,Fourier transform (FT)-Raman spectroscopy has been reintroduced to deal
with the problem of strongly fluorescing organic compounds?1p32 The instrument,
illustrated in Figure 2, uses the 1.064 pm near-infrared line of a Nd:YAG laser for

LN2 Cooled
Ge detector Fixed
mlrror

line filter

Reproduced with permission of the American Chemical Society.

Figure 2 Optical setup for the FT Raman spectrometer?*


View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 117

excitation of the sample (180’ backscattering excitation/collection geometry) in order to


avoid exciting the transitions that induce fluorescence in many organic, polymeric, and
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

biological compounds (usually excited by greenblue light). Unfortunately, the Nd:YAG-


induced Raman signal is too weak to be measured by conventional dispersive instruments
as expected due to its v4 frequency dependence. This problem is solved, however, by
multiplexing the signal using a Michelson interferometer to obtain both the multiplex and
throughput advantages (Fellgett and Jacquinot advantages, respectively). The quasi-
elastically scattered stray-light components must be removed by an optical filter system
before the signal reaches the near-IR detector (Ge, liquid-N2 cooled). This is because the
Downloaded by McGill University on 26 December 2012

high level of random noise present in the very intense laser line would otherwise be
distributed throughout the entire Raman spectrum by the R-distribution process, thereby
degrading the entire spectrum.
FT Raman has been shown to be very promising in the routine acquisition of
Raman spectra of fluorescing organic compounds. The technique has been used to collect
the Raman spectra of strongly fluorescing compounds such as anthracene, Rhodamine-6G
(laser dye), cyclohexane, phenol, benzene/toluene mixtures, as well as biologically active
macromolecules~3 The use of FT Raman has also spread to surface-enhanced studies of
~ ~ ~area of heterogeneous catalysis,
pyridine at silver, copper, and gold e l e ~ t r o d e s . 3In~ the
FT Raman could be a valuable asset in the analysis of fluorescing organic products and
reactants in the study of chemical kinetics and mechanisms. Of more recent interest is the
development of an FT-Raman microprobe in~trument.~~
The Hadamard-transform (HT) technique has recently emerged as a viable
alternative to FI’methods for recording spectra from low intensity signals or at high speed
or high s e n ~ i t i v i t y . 3This
~ ~ ~revival in HT techniques is a consequence of the convention-
al, mechanically operated HT encoding masks (multiple-slit masks which are mechanically
moved to allow different segments or resolution elements of the spectrum through to the
detector at any one time) being replaced by new electro-optic arrays as a result of
technological advances in the development of electro-optic switching arrays. Generally
speaking, €IT-spectrometry involves the application of multiplex spectrometry to
dispersion spectrometry.
A prototype HT spectrometer has been designed to cany out Raman spectroscopic
in~estigations.3’~~The spectrometer resembles that of a Czemy-Turner scanning
spectrometer possessing two 118O-grooves/mm gratings. The first grating serves to
disperse the radiation onto a 127-element encoding mask (127 slits-liquid crystal spatial
light modulator; spectral range = 300-2500 nm) where it is Hadamard modulated (about
half the slits are programmed to be electro-optically opened at any one time) and the
second grating dedisperses the remaining resolution elements of light. Those spectral
elements that are allowed to exit the encoding mask impinge upon a single detector, and
the measured intensity is equal to the additive contribution of the open spectral elements
of the mask. After N numbers of slit arrangements have been sampled, and after N
intensities have been measured for these N arrangements, the N linear equations are
View Online
118 Catalysis

simultaneously solved for the intensities of each resolution element of the Raman
spectrum. HT multiplexing of the Raman signal has been shown to yield spectra faster
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

than the conventional scanning dispersive instrument. This capability allows the study of
dynamic processes in time (real-time analysis) comparable to that of an optical
multichannel detector without incurring the significant expense of the photodiode array.
HT Raman offers itself as a viable replacement for FT-Raman instruments and CCD-based
instruments in the near-infrared region because of its intrinsic simplicity in design and the
ease with which the laser line is readily masked from the detector (i.e., via a knife-edge
or notch rejection filter). More recent developments include HT-Raman microscopy for
Downloaded by McGill University on 26 December 2012

profile imaging, optical sectioning, and multispectral imaging for multicomponent


materials?84 To date, commercial HT-Raman instruments are not yet available.

4.6 Colloidal Filters. - State-of-the-art Raman instrumentation has progressed at a


remarkable rate in recent years, especially with regard to advances in solid-state detectors;
these advances, in particular, have led to very sensitive IPDA and CCD detectors. Raman
spectrometers have also improved significantly in design, but the problem of stray light
is still sometimes present even in triple monochromator systems. A crystalline colloidal
filter has been developed which effectively eliminates much of the stray light of the laser
beam from Raman spectra. Impressive rejection capabilities have been observed for a 1/3
meter single-spectrograph stage preceded by the crystalline colloidal filter!1 This filter
consists of crystalline array of polystyrene spheres ( 4 . 1 pm) with lattice spacing of a few
hundred nanometers which is comparable to the visible wavelength of the excitation
source. The laser frequency is Bragg diffracted so that its transmission through the filter
is less than lo-''. The Raman-shifted bands are transmitted with greater than 70%
efficiency for the Stokes lines and 60% for the anti-Stokes lines. Furthermore, the
crystalline colloidal fdter selectively rejects the quasi-elastically scattered components so
as to allow the simultaneous collection of both the Stokes and the anti-Stokes components
without any mechanical movement of the spectrograph (withOMA detection). The major
advantage of the colloidal filter is that the stray-light rejection capability of the triple
monochromator is no longer required for OMA detection. Instead, a single-stage
spectrograph, which is a much less expensive alternative and has a much greater
throughput, can be employed. The crystalline colloidal filter is patented by Prof. Asher
(University of Pittsburgh)!'

5 SamDlinn Techniaues

5.1 SamDle PreDaration and Handling. - There are a wide variety of sampling techniques
used for collecting Raman spectra of solid catalyst systems depending on the type of
sample and the conditions under which the measurement is to be made. Simple sample
holders can be used if the sample is stable under ambient and atmospheric conditions.
ThC sample morphology is not as important in Raman spectroscopy as it is in other types
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 119

of spectroscopy, such as infrared spectroscopy, since the sample need not be transparent
or particularly smooth. Nevertheless, powdered samples are typically pressed into the
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

form of opaque, self-supporting wafers for greater ease and consistency in sample
manipulation. If only milligram quantities of sample are available, a KBr backing is used
to support the small quantity of material and the two are pressed together as one self-
supporting body.
For colored samples, a significant amount of the incident radiation may be absorbed
and cause local heating; the dramatic local temperature rise of the sample may change or
degrade the sample considerably. This problem is most often encountered in resonance
Downloaded by McGill University on 26 December 2012

Raman studies where the frequency of the laser is purposely chosen to overlap with the
electronic absorption maximum of the sample. To avoid the problem of local heating, the
incident beam is defocused over a greater area of the sample. This decreases the incident
flux of light onto the sample and the associated temperature rise is reduced. Alternatively,
the sample may be cooled. The most common method used for preventing a temperature
rise in solid samples (crystals and pressed powders) involves rotating the sample at speeds
of about 2000 rpm so as to allow the heat induced by absorption to be conducted away
from the probed area before the laser reilluminates the same area after each successive
revolution. These sample rotators are commercially available from Raman instrument
manufacturers.

5.2 In Situ Investigations.- Raman spectroscopy is ideally suited for the in situ investiga-
tion of catalysts because there are no stringent requirements on sample preparation or on
the sample environment within the cell. Many considerations must be taken into account,
however, when designing an in situ Raman cell for catalytic studies. First, the cell must
be able to adequately accommodate the sample, and the environment around the cell must
be capable of being precisely controlled and monitored. Secondly, the in situ cell must
be spatially and optically coupled to the Raman spectrometer. Often, the optics of the
Raman instrument must be altered in some way (e.g., collection optics) in order to
accommodate the cell. The sample and cell should be designed and positioned to
minimize the extraneous Raman scattering arising from the cell walls and windows.
Precise optical x-y-z translation stages, tilting, and rotation stages may be used to move
the cell accurately and reproducibly with respect to the optical axis and the acceptance
angle of the monochromator.
Several types of cells designed for in situ Raman studies have appeared in the
literature for applications in heterogeneous catalysis. For example, the rotating cell of
Cheng et alp2 has been used for controlled atmosphere studies and for collecting in situ
Raman spectra of pyridine adsorbed on supported molybdate catalysts at temperatures of
600 OC and pressures as low as 1.333 X lo4 Pa. Chan et al!3 developed an in situ
Raman cell for heterogeneous catalysts, shown in Figure 3, capable of sample rotation and
heating of the sample in air up to 600 OC. The sample holder was made from a corrosion-
resistant alloy, and the catalyst disc held beneath a collar that screws onto the holder. The
View Online
120 Catalysis

Ouartr Shield
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

ser Heater Thrrmncnunlr


Downloaded by McGill University on 26 December 2012

Figure 3 Schematic of in sltu cell arrangement.

sample holder was coupled to a rotatable shaft and driven by an air motor for rotation.
Reaction gas mixtures were introduced into the cell from a manifold at a rate of 50 to 500
cm3/min with a delivery pressure of 20 to 30 kPa. Heating was accomplished with a
cylindrical heater coil surrounding the quartz cell. There are many different variations on
sample cells that have been used in conditions of extreme temperatures and pressure as
well as cell designs for liquids, gases, and special applications. These are described in
laboratory guides and methods on Raman ~pectroscopy.~

5.3 Samde Fluorescence. - The problem of how to cope with a fluorescence background
has been one of the pervading issues in Raman spectroscopy of industrially important
systems. The characterization of a catalytically important surface species present on a
substrate, for example, is often complicated by the presence of strongly fluorescing
impurities arising from either the substrate or, worse, the species itself. A variety of
methods have been proposed to deal with the problem of fluorescencesuch as instrumental
approaches including FT-Raman, time-resolvedspectroscopy,nonlinear Raman techniques,
and simply by altering the wavelength of the excitation source so as to decrease the
probability of pumping the fluorescing transition. Besides instrumental approaches, the
Raman spectrum can also be mathematically enhanced, for example, using the maximum
entropy method (MEM) to enhance the Raman data over the fluorescence by fdtering out
random noise from structured e~ents.4~
Other methods of reducing the fluorescence background from the Raman spectrum
focus on the sample itself. The easiest method, and perhaps the one requiring the greatest
degree of hand waving, is the so-called "drench-quench" technique which involves a
prolonged exposure (minutes to hours) of the sample to the full power of the laser beam
in order to "bleach out" the fluorescing centers uresent in the samde matrix. The most
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 121

universal method of reducing the fluorescence background, however, is the chemical


approach which involves taking the necessary steps to "clean up" the sample. Supported
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

metal-oxide systems, for example, are generally calcined in air at 500 to 550 OC prior to
collecting Raman data to remove any fluorescing species intruding from the environment
such as polyaromatic hydrocarbons (PAHs).

6 ADDlications of Raman Spectroscopv to Heterogeneous Catalvsis

Numerous publications have appeared during the past several years describing
Downloaded by McGill University on 26 December 2012

applications of Raman spectroscopy to catalysis; these are cited in several excellent


reviews (see references 2-4; 46-49). Most of the reported investigations were performed
with instrumentation that is inferior to the current technological standards of Raman
spectroscopy. Thus, the purpose of this section is to illustrate what is currently achievable
in Raman spectroscopy of heterogeneous catalysis by selecting a few recent key examples
from the literature.

6.1 Bulk Mixed Oxide Catalvsts. - Raman spectroscopy of bulk transition metal oxides
encompasses a vast and well-established area of knowledge. The fundamental vibrational
modes for many of the transitional metal oxide complexes have already been assigned and
tabulated for systems in the solid and solution phases." Perhaps the most well-known and
established of the metal oxides are the tungsten and molybdenum oxides because of their
excellent Raman signals and applications in hydrotreating and oxidation catalysis.
Examples of these two very important metal-oxide systems are presented below for bulk
bismuth molybdate catalysts, in this section, and surface (two-dimensional) tungstate
species in a later section.
6.1.1 Molvbdate Reference ComDounds. In Raman spectroscopy, characteristic
vibrational bands derived from bulk reference compounds have been used to distinguish
between tetrahedral and octahedral transition metal-oxide species in catalytically important
rnaterial~.'~~ For
' ~ molybdates,
~~~~ the Raman spectra of BqXvIoO, (octahedral),
MoO;-(aq) (tetrahedral), and molten K2M0207(tetrahedral dimer) may be used to show
typical spectral differences between octahedral, tetrahedral, and tetrahedral dimeric
molybdate species. Stick diagrams representing the Raman spectra of these reference
compounds are presented in Figure 4. Ba@MoO, is an ordered perovskite consisting
of a slightly distorted MOO, octahedron. For a perfect MOO, octahedron, three Raman
bands are expected to be observed according to the 0, point-group character table: r,,
= vl(Al,) + v2(Eg)+ ~3(E2,)(see Section 3.1); as the octahedron becomes more distorted,
a progressively greater number of the 15 possible vibrational modes become Raman-
activated (3N- 6 = 3(7) - 6 = 15). If the Moos octahedron in B a + M o O , were assumed
to be of perfect 0,point-group symmetry, then the vl(Al,) symmetric Mo-0 stretching
mode may be assigned to the band observed at 812 cm-', the v2(Eg) symmetric stretch to
the band at 650 cm-', and the v3(F2,) bending mode to the band at 416 cm-lF3 A
View Online
122 Catalysis
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

1200 1000 800 600 400 200


Raman Shift (cm-l)
Flgure 4 Stick dlagrams representlng the Raman spectra of molybdenum oxlde reference
compounds.

comparison of the observed Mo-0 stretch at 812 cm" with that predicted for the perfect
MOO, octahedron at 665 cm-' (by the FLS approach; see also Table 2), however, shows
a significant discrepancy of 147 cm-', thereby placing doubt on the assumption of perfect
0, point-group symmetry for the MOO, octahedron in B@hMo06.
A perfect MOO, tetrahedron, having Td point-group symmetry, is expected to
exhibit four fundamental modes of vibration, rvib, = vl(Al) + v2(E) + v3(F2)+ ~4O;2),as
discussed earlier (see Section 3.1). If the aqueous molybdate ion in basic solution,
MoO:-(aq), were assumed to be a perfect tetrahedron, then its highest occurring Raman
band at 897 cm-' may be assigned to the totally symmetric vl(Al) stretching mode, the
less intense band at 841 cm-' to the antisymmetric ~3(F2)stretching mode, and the band
at 318 cm-' represents the degenerate symmetric v,(E) and the antisymmetric vq(F2)
bending modes.52 A comparison of the highest occurring Mo-0 Raman stretching
frequency at 897 cm-' with that estimated for the perfect MOO, tetrahedron frequency at
858 cm-' (Table 2) shows that the MoO:-(aq) species does not possess perfect Td point-
group symmetry, but is slightly distorted. In fact, the distorted MOO, structures in
CaMoO, (880 cm-') and B-Bi2M0209(887 cm-') are more regular than MoO?(aq) as
evidenced by their lower Mo-0 stretching frequencies, and the MOO, structures in
N%Mo0,-2H20(897 cm-') and Bi3(Fe04)(Mo04)2(897 cm-') are at least as regular as
MoOz-(aq) because the Mo-0 stretching frequencies are identical.', This analysis leads
one to consider alternative structural units for the molybdate species in MoOz-(aq); for
example, speculation leads to a partially protonated MOO,-type (aq) species (perhaps of
C,, point-group symmetry) that is consistent with the two observed Mo-0 stretches at 897
(two nonbridging or terminal Mo-0 bonds) and 841 cm-' (two Mo-O"-H bonds).
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 123

The tetrahedral dimeric species K2M0207possesses Raman features similar to those


of the tetrahedral monomer MoOz-(aq) because the nature of the distortion on the
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

tetrahedra is similar. The differences observed in their Raman spectra may be related to
the structural differences between the tetrahedral MOO, monomer and Mo207 dimer.
Whereas there is only one type of Mo-0 bond present in the perfect MOO, tetrahedron
(nonbridging for the isolated species), there are two types of Mo-0 bonds present in the
Mo207 dimer: bridging and nonbridging. The bridging bond is expected to possess
roughly one-half the bond strength of the nonbridging bonds, resulting in a much lower
Mo-0 stretching frequency. In order to preserve the valence of the Mo6+ cation (Csi =
Downloaded by McGill University on 26 December 2012

6.00 vu), the presence of the bridging bond would result in an increase in the bond
strength and stretching frequency of the nonbridging Mo-0 bonds. The stretching
frequency of the nonbridging Mo-0 bonds observed at 927 cm-' corresponds to a bond
strength of 1.707 vu (see Table 1); this leaves a bond strength of 0.880 vu for the bridging
Mo-0 bonds and corresponds to a calculated stretching frequency of 61 1 cm-' (using eqs.
and parameters from Table 1, and the valence sum rule). The estimated value at 61 1 cm-'
for the stretching of the bridging Mo-0 bond in the Mo207 dimer is very close to that
observed as a very weak band at 565 cm". Thus, based on Gv site symmetry for the
Mo6+cation in molten K2M0207,the 927 and 883-cm-' bands are assigned to symmetric
and antisymmetric stretching modes, the 565-cm-'band to the symmetric stretching of the
bridging Mo-0 bonds, the band at 350 cm-' to the bending of the nonbridging O=Mo=O
functionalities, and the band at 196 cm-' to the bending of the Mo-0-Mo linl~age.5~
6.1.2 Bismuth Molvbdate Catalvsts. The Raman spectra of the bismuth molybdates, with
Bi/Mo stoichiometric ratios between 0.67 and 14, have been examined using the F'LS
approach (see Section 3.2)?, The bismuth molybdates fall into an unusual class of
compounds, the ternary bismuth oxide systems Bi-M-0 (where M = Mo, W, V, Nb, and
Ta) which exhibit a variety of interesting physical and chemical properties. Of
commercial importance, the bismuth molybdates are heterogeneous catalysts for selective
oxidations and ammoxidations (the Sohio process), for example, propylene (C3H6) to
acrolein (%H40) by oxidation or to acrylonitrile (C3H3N)by ammoxidation?s~s6
The observed Raman bands for the bismuth molybdate samples (0.67 I Bi/Mo I
14) have been tabulateds4 and show that many of the compositions examined are
multiphasic. As expected, the bismuth molybdate and bismuth oxide phases were also
found to depend on the Bi/Mo ratio, with the following phases being identified by Raman
spectroscopy: a-Bi2Mo3OI2, B-Bi2M0209,y-Bi2MO06, y'-Bi,MOO, (high-temperature
phase), Bi&fO,O1,, Bi37M07078,a Bi-Mo-0 sillenite phase (y-Bi,O, structure), a-Bi203,
and metal-cation stabilized 6- and &phases of Bi203. The complete crystal structures of
a-Bi,Mo,O,,, B-Bi,Mo,O, and y-Bi2MO06 have been reported in the literature. The a-
Bi,Mo30,, phase is reported to consist of three crystallographically distinct Moos units:7
B-Bi,Mo20, to consist of four distinct MOO, tetrahedra with similar ranges of bond
lengths:8 and y-Bi2Mo06 has a slightly distorted MOO, ~ c t a h e d r o n . ~Each
~ of the
View Online
124 Curulysk

reported Mo-0 bond lengths in each of these phases has been assigned to its correspond-
ing observed Mo-0 Raman stretching frequency by utilizing the FLS relations.54
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

In spite of the numerous electron diffraction studies carried out on higher order
bismuth molybdates, consisting of y'Bi2MO06 (high-temperature form of y-Bi2MOO6),
Bi6Mo,0,,, Bi37M07078,and the sillenitephase, the molecular structures of the molybdate
species in these systems were left This was probably due to the
multiphasic nature of these materials, their exceedingly large superlattices,and the inherent
difficulty associated with determining the positions of oxygen atoms using electron
diffraction techniques.
Downloaded by McGill University on 26 December 2012

The Raman spectra for the bismuth molybdate samples with Bi/Mo ratios of 2:1,
3:1, 4:1, 38:7, and 14:l are shown in Figures 5 through 7. The major phase present in
the high-temperature 2:l sample is y'-Bi,M006, Figure 5 , while that present in the 3:l
sample is Bi,MO,O,,, in the 4:l and 38:7 samples Bi37M07078is the major phase, Figure
6, and the Bi-Mo-0 sillenite phase is the major phase at the 14:l composition, Figure 7.

097
I

0 200 400 600 800 1000 1200 1400


Raman Shift (cm")
Figure 5 Raman spectrum of the high-temperature form of BI,O,-MoO,, showing yc
Bi,MoO, as the maJorphase and the BI-Mo-0 slilenlte phase as the minor component
(dashed curve).
The structure determination of the molybdates species in fBi2MOO6, using the
FLS approach, is shown in Table 3. Six Mo-0 stretching frequencies are observed in the
Raman spectrum, from Figure 5 , at 897, 865, 829, 822, 788, and 771 cm-'. Each Mo-0
stretching frequency is initially assumed to be due to the localized stretch of a unique Mo-
0 bond type. The FLS equations and parameters, Table 1, are used to convert these
Raman stretching frequencies to bond lengths and Pauling bond strengths which are
tabulated at the top of Table 3. Next, the bond strengths are added in every possible
combination to see, mathematically, which combinations yield the formal oxidation state
of the M Ocation,
~ which is 6.0 vu, within a specified tolerance. As Table 3 shows, there
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 125
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

0 200 400 600 800 1000 1200 1400


Raman Shift (cm-')

Flgure 6 Raman spectrum of (a) BkMo = 3:l with BI,Mo O,, as the major component
(bottom), (b) BI:Mo = 4:l (mlddle), and (c) BI:Mo = 38:f (top) with BI,MO,O,~ as the
major component.

I 143 I

0 200 400 600 800 1000 1200 1400


Raman Shift (cm-l)

Figure 7 Raman spectrum of 7B1203-Mo0, (BI:Mo z 14:l) showing the BI-Mo-0 slllenlte
phase as the maJorcomponent and a-B1203 as the mlnor component (dashed curve).

are 20 possible molybdate structures, labeled (a) through (t). It is interesting to note that
all possibilities shown in Table 3 have coordination numbers of four; that is, all are MOO,
tetrahedra. There must necessarily be more than one type of MOO, species, however,
because there are more than four different Mo-0 stretching modes representing more than
four different Mo-0 bond types. If two distinct MOO, species were present, then the
View Online
126 Catalysis

Table 3 y’-Bi,MoO,, Structure determination by Raman spectroscopy


Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

Mo-0 bond types A 897 1.615 1.738


B 865 1.520 1.755
C 829 1.418 1.776
D 822 1.398 1.780
E 788 1.306 1.800
Downloaded by McGill University on 26 December 2012

F 77 1 1.261 1.811
Possible structures
Coord. no. A B C D E F Valence (XsJ
4 2 0 0 1 0 1 5.889
4 3 0 0 0 0 1 6.105
4 2 1 0 0 0 1 6.01 1
4 2 0 1 0 0 1 5.908
4 1 2 0 0 0 1 5.916
4 2 0 0 1 1 0 5.934
4 2 1 0 0 1 0 6.055
4 2 0 1 0 1 0 5.953
4 2 0 1 1 0 0 6.045
4 2 0 2 0 0 0 6.065
4 2 0 0 2 0 0 6.026
4 1 2 0 0 1 0 5.961
4 1 2 0 1 0 0 6.053
4 1 2 1 0 0 0 6.073
4 1 1 1 1 0 0 5.951
4 1 1 2 0 0 0 5.970
4 1 1 0 2 0 0 5.932
4 0 3 0 1 0 0 5.959
4 0 4 0 0 0 0 6.081
4 0 3 1 0 0 0 5.978
Best structures
hybrid (af) (0
2xA 2~1.738(16) 8, 3xB 3~1.755(16) A
1xD 1~1.780(16) 1xc 1~1.776(16)
lxEF 1x1305 (21)
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 127

shortest Mo-0 bonds in these two structures would give rise to two Mo-0 stretches above
858 cm-' (estimated for the perfect MOO, tetrahedron [see Table 21 - all distorted MOO,
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

species must stretch above 858 cm-'). As the simplest case, the two Mo-0 stretching
bands at 897 and 865 cm" are assumed to represent the stretching of the shortest Mo-0
bonds in two structurally different MOO, tetrahedra in the y'Bi2MOO6 phase. The final
step is to find two MOO, species, from the structural possibilities listed in Table 3, for
which their Raman bands would give rise to the observed Raman spectrum (that is, for
which all Raman stretching frequencies are accounted for). The best combination of
structural possibilities consistent with the Raman spectrum in Figure 5 is a hybrid structure
Downloaded by McGill University on 26 December 2012

of (a) and ( f ) (using bonds A, D, E, and F), and structure (t) (using bonds B and C). The
former MOO, species has bond lengths of 2 x 1.74(2), 1 x 1.78(2)and 1 x 1.80(2)A, and
the latter species is very regular with bond lengths of 3 x 1.76(2) and 1 x 1.78 (2) A;
these results are summarized at the bottom of Table 3. The two MOO, tetrahedra found
to be present in y'-Bi,MOO6 compare favorably with the model postulated by Buttery,
using electron diffraction, which suggested the presence of two types of MOO, species in
the y'Bi2MOO6 structure: isolated molybdate tetrahedra forming square tunnels and
enclosing a single column of Bi3+ sites, and a second tetrahedral species linking pairs of
these tunnels.60
The structure determinations of the molybdate species in Bi6M02O15, Bi37M07078,
and the sillenite phase using the FLS approach are shown in Tables 4 and 5. The Raman
spectrum of the Bi6MO@,, phase, shown as the bottom spectrum in Figure 6, exhibits
Mo-0 Raman stretching modes at 880 and 800 cm-', and that of Bi38M07078shows
almost identical bands at 870 and 800 crn-'. Table 4 shows that the FLS structural
determination is trivial with only one set of Raman bands, or one possible structure, for
each phase. Both the Bi6Mo2Ol5and Bi38M07078 phases possess MOO, tetrahedra that
are identical within experimental error with bond lengths of 3 x 1.75(2)and 1 x 1.79(2) A.
The Raman spectrum of the Bi-Mo-0 sillenite phase, shown in Figure 7 (all bands below
800 cm-' are due to bismuth oxide intemal/extemal modes of the y-Bi203-typestructure
and the a-Bi203 co-existing phases), shows Mo-0 stretches at 877, 820, 815 and
800 cm-'. Table 5 shows that there are five possible molybdate structures and that all are
MOO, tetrahedra. All Mo-0 stretches are utilized by forming a hybrid of structures (a)
and (e), and this results in Mo-0 bond lengths of 2 x 1.75(2), 1 x 1.76(3) and 1 x
1.79(2) A. By analogy to molybdate reference compounds, the MOO, tetrahedron in the
Bi-Mo-0 sillenite structure is similar to that in CaMoO, which has reported Mo-0 bond
lengths of 1.775 AM and Raman stretching modes at 879 (most intense), 848 and
794 ~m-'.~'
In addition to the capability of Raman spectroscopy in determiningthe coordination
numbers and bond lengths of molybdate species in bismuth molybdate phases, Raman
spectroscopy can also be used as an in situ probe for the bismuth molybdates under
reaction conditions. In situ Raman studies have been carried out, for example, on the 6-
Bi2Mo,09 phase under redox conditions, where insights into the surface mechanism of the
View Online
128 Catalysis

Table 4 Bi,Mo,015 and Bi,Mo707,, Structure determination by Raman


Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

Spect-PY
B&Mo2015 v (cm-') s (vu) R (A)
Mo-0 bond types A 880 1.564 1.747
B 800 1.338 1.793
Possible structure
Coord. no. A B Valence (ZsJ
Downloaded by McGill University on 26 December 2012

(a) 4 3 1 6.030
Best structure for Bi,Mo,015
(a)
3xA 3~1.747(16) A
1XB 1~1.793(16)
~~~~ ~

Bi38M07078 v (cm-'1 s (vu) R (4


Mo-0 bond types A 870 1.535 1.752
B 800 1.338 1.793
Possible structures
Coord. no. A B Valence @si)
(a) 4 3 1 5.942
Best structure for Bi3,M~07,
(a)
3xA 3~1.752(16) A
1XB lx1.793 (16)

conversion of propylene to acrylonitrile (ammoxidation) have been realized.66 These


mechanisticinsights include direct spectroscopic evidence for solid-state disproportionation
of the metastable B-Bi,Mo,O, phase under redox conditions, the identification of key
catalytic phases in bismuth-iron molybdate systems, and the direct identification of
catalytic function of active oxide ions in bismuth molybdate selective oxidation catalysts.
In situ Raman spectra were also examined for y-Bi,MoO, reduced with either 1-butene,
propene, methanol, or ammonia and reoxidized with These experiments
revealed a multifunctional active site for propene oxidation and ammoxidation, in which
a-hydrogen abstraction occurs by oxide ions bridging bismuth and molybdenum atoms.
Furthermore, 0, chemisorption, reduction, and dissociation were attributed to the directed
lone pairs of electrons associated with Bi-0-Bi linkages in the structure.
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 129

Table 5 Structure determination of the molybdate species in the Bi-Mo-0 sillenite


Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

phase by Raman spectroscopy

~~ ~

Mo-0 bond types A 877 1.555 1.748


B 820 1.393 1.781
C 8 15 1.379 1.784
D 800 1.338 1.793
Downloaded by McGill University on 26 December 2012

Possible structures
Coord. no. A B C D Valence @si)
(a) 4 3 0 0 1 6.004
(b) 4 3 1 0 0 6.058
(c) 4 3 0 1 0 6.045
(4 4 2 2 0 0 5.896
(el 4 2 1 1 0 5.882
Best structure
hvbrid (ae)
2xA 2~1.748(16) A
1xAB 1~1.765(32)
lxCD 1~1.788(20)

e d Oxide Catalvsts. - Supported metal oxide catalysts are formed when


6.2 S u ~ ~ o r t Metal
one metal oxide component is deposited onto a second metal oxide substrate such as
A1,0,, TiO,, or SO,, to name just a few of the support materials used in heterogeneous
catalysis. Figure 8 illustrates how a supported metal-oxide may be present as agglom-
erated metal-oxide crystallites, a two-dimensional metal-oxide overlayer (surface metal-
oxide phase), or a bulk mixed-oxide phase with the oxide substrate. The state of the
supported metal oxide component depends on the nature of the metal-oxide/support
interaction and the calcination temperature. The scientific development of this area of
catalysis has progressed very slowly because usually several of the above metal-oxide
phases are "simultaneously" present in supported metal-oxide catalysts, and more
conventional characterization methods are unable to discriminate between the various
metal oxide phases. Raman spectroscopy, however, is uniquely suited for the characteriza-
tion of supported metal oxide catalysts because it is capable of discriminating between the
various phases of the supported metal oxide component as well as providing molecular
level structural information.
View Online
130 Catalysis
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

Oxide
Support
Downloaded by McGill University on 26 December 2012

Agglomerated Metal Two-Dlmenslonal Metal Bulk Mixed Oxlde


Oxide Crystallltes Oxide Overlayer Phase
(Surface Metal Oxlde (M = Metal Oxlde)
Phase) (S = Oxide Support)

Figure 8 The possible states of a supported metal oxlde phase.

6.2.1 Sumorted Metal Oxides as Surface Oxides. The ability of Raman spectroscopy to
discriminate among the various metal-oxide phases present in supported metal oxide
catalysts is demonstrated by recent studies on the WO,/AI,O, ~ y s t e m . ~ ' * ~The
* - ~Raman
'
spectra of crystalline WO,, crystalline AI,(WO,),, and 10% WO, supported on y-A1203
are presented in Figure 9. Crystalline WO, possesses numerous Raman-active vibrational
modes with major Raman bands at 808, 711, and 273 cm-'.68 Crystalline A12(W04),,
formed by the solid-state reaction between tungsten oxide and alumina at elevated
temperatures, exhibits strong bands at 1057, 386, and 371 cm-'. The 10% W03/A120,
sample containing the supported tungsten-oxide phase does not exhibit the strong Raman
bands characteristic of crystalline WO, or AI,(WO,),, but has weak broad bands at 986,
870, and 300 cm-'. The y-Al,O, support does not exhibit any Raman active modes. The
relative Raman cross-sections of these tungsten oxide phases are also very different. The
crystalline WO, band at 808 cm-' is approximately 160 times stronger than the supported
tungsten oxide band at 986 cm-' per tungsten oxide, and the crystalline AI,(WO,), Raman
band at 1057 cm-I is approximately five times stronger than the supported tungsten oxide
band per tungsten oxide. Furthermore, the Raman band position of the supported tungsten
oxide exhibits a dependence on tungsten oxide surface coverage and shifts from 950 to
lo00 cm-' due to structural changes in the supported tungsten oxide The
Raman bands of the crystalline WO, and A12(W04), phases, however, are fmed in
position and do not shift with crystalline WO, and AI,(WO,), content. The Raman
vibrational spectra reveal that the state of the supported tungsten oxide on alumina is very
different from that found in crystalline WO, and A12(W04),. With the assistance of in
situ Raman investigation^!^ to be discussed below, it was conclusively demonstrated that
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 131

IJL 10% WO$AI,O, (500 "C) -


Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

1200 1000 800 600 400 200


Raman Shift (cm-I)
Figure 9 Raman spectra of crystalline WO,, crystalline AI,(WO,),, and 10% WOdAI,O,
(calclned at 500 "C).

the Raman bands of the supported tungsten-oxide phase originate from a two-dimensional
tungsten oxide overlayer on the alumina support. For an alumina support possessing 180
m2/g, approximately 25% to 30% WO, corresponds to one complete monolayer of the
surface tungsten-oxide phase, and tungsten oxide in excess of this amount forms
crystalline WO, particles.
6.2.2 In Situ Raman S m t r o s c o ~ vof Surface Metal Oxides. Another advantage of the
Raman technique is that experiments can be performed in situ, under conditions in which
the sample environment (temperature and gas-phase composition) can be controlled. Such
experiments provide additional information about the supported metal-oxide phases as
shown in Figure 10 for 7% WO, dispersed on a TiO, support. The Raman spectrum of
the sample held at room temperature, Figure lqa), exhibits a broad band at 970 cm-' that
is characteristic of supported tungsten oxide. Heating the WO,PiO, sample to 500 "C
in flowing dry-air results in a simultaneous sharpening and shift of the supported tungsten-
oxide Raman band to a higher wavenumber, Figure lqb). These changes are due to the
desorption of moisture originally present on the sample69and reveal that water molecules
directly coordinate to the supported tungsten-oxide phase and affect its state. Cooling the
sample to 80 O C , Figure lqc), causes a slight sharpening of the supported tungsten-oxide
Raman band because of the removal of thermally induced band broadening. The original,
broad band at 970 cm-' of the supported tungsten oxide can only be restored after the
sample is exposed to moisture, Figure lqd). Similar in situ Raman studies with
supported crystalline metal-oxide phases demonstrated that the band positions of
crystallites are independent of temperature and the presence or absence of moisture.69
View Online
132 Catalysis

-
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

10

p
f -
1
E

6 -
Downloaded by McGill University on 26 December 2012

4 -

2 -

800 900 1000 1100 12M) 1m


Wavenumber (cm-1)
Figure 10 In sltu Raman spectra for 7% WOgO,. Sample at temperatures of: (a) 20 "C,
(b) 500 "C, (c) 80 "C, and (d) 20 "C ambient envlronment.

The ability to perturb the Raman spectrum of tungsten oxide dispersed on TiO, by
the adsorption/desorption of water molecules reveals that the supported tungsten oxide is
not present as a three-dimensional crystalline/amorphous phase or incorporated into the
titania-supported structure but must be present as a two-dimensional surface metal oxide
overlayer "adsorbed" on the titania support. This is illustrated schematically in Figure 11.
Complementary infrared studies show that the surface metal-oxide overlayer coordinates
to the oxide support by reacting with the surface hydroxyl groups.72 Similar in situ
Raman behavior is also observed for other metal-oxide systems (rhenium oxide, chromium
oxide, molybdenum oxide, tungsten oxide, niobium oxide, and vanadium oxide) on
alumina, titania, and silica supports12713969773-85
demonstrating that these metal oxides are
also adsorbed on the surface of the oxide supports.
The Raman spectra of crystalline NiO, crystalline NiA1,0,, and 8% NiO supported
on y-Al,O, are presented in Figure 12. Materials containing nickel oxide possess Raman
signals that are about 100 times weaker than the corresponding tungsten-oxide materials
and occur at 500 to 600 cm-' in comparison to 800 to 1100 cm-' for the tungsten-oxide
materials. Both of these effects originate from the much lower Ni-0 bond order present
in the nickel-oxide materials. Crystalline NiO exhibits Raman bands at 460 and 500 cm-',
and crystalline NiAl,O,, formed by the solid-state reaction between NiO and Al,O, at
elevated temperatures, exhibits major Raman bands at 375 and 600 cm-'. The 8%
NiO/A1203 sample containing the supported nickel oxide phase does not exhibit the
Raman bands of crystalline NiO or NiA1,0, but has a very broad band at 550 cm-'. The
Raman vibrational spectra demonstrate, as discussed earlier for the supported tungsten-
oxide system, that the state of the supported nickel-oxide system on alumina is very
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 133
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

H\0 wox
'H [
8 H
'
-H20)
Downloaded by McGill University on 26 December 2012

Figure 11 Schematic diagram showing the effect of adsorption/desorption of moisture on


a tungsten oxide-titanla system.

36 -
h

6a 32-

E
3 28
-
0
(3
g
v
21
-
f 20-
Q)
CI

I6 -
I
g
K
12-

8 -

4 -

100 200 300 400 500 600 mo 800 goo

V(Cm-1)

Figure 12 Raman spectra of crystalllne NIO, crystalllne NIAI,O,, and 8Oh NIO/AI,O,.

different from that found in crystalline NiO or NiAl,O,. Additional information about the
state of the supported nickel oxide on alumina is provided by in situ Raman studies.
The response of the supported nickel oxide on alumina to an in situ experiment is
very different from that presented earlier for the supported tungsten oxide, as shown in
Figure 13. The Raman band of the supported nickel oxide on alumina is not affected by
the removal of the moisture present on the sample surface (compare Figures 13(a) and
13(d)) and suggests that water molecules do not coordinate to the supported nickel oxide.
View Online
134 Catalysis

40
1
13% NIOIA120 (450'C)
In Sttu Caktinatlon I
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

(d) 20 'C
Downloaded by McGill University on 26 December 2012

100 200 300 400 500 600 700 800 900 1000

v(crn-1)

Flgure 13 In situ Raman spectra for 13% NIO/A120,. Temperatures: (a) 20 "C, (b) 400 "C,
(c) 600 "C for 2.5 hr, and (d) 20 "C.

The pronounced thermal broadening of the supported nickel-oxide Raman band at elevated
temperature (Figures 13(b) and 13(c)) suggests that the nickel-oxide vibrational modes
may be intimately coupled to the vibrations of the alumina supp0rt.8~ The water
insensitivity of the supported nickel-oxide Raman band and its pronounced thermal
broadening at elevated temperatures are consistent with the incorporation of the nickel
oxide into the y-Al,O, surface (formation of a surface spinel). This suggests that nickel
oxide is not adsorbed on the alumina support as is surface tungsten oxide, but that the
surface nickel oxide is "absorbed" into the surface layer of the alumina support. This is
presented schematically in Figure 14. Thus, in situ Raman spectroscopy studies are
capable of distinguishing between metal oxides that are adsorbed on the alumina support
surface and metal oxides that appear to be absorbed into the alumina support surface as
a surface spinel. Similar Raman characteristics are also observed for other metal oxides
(cobalt oxide, copper oxide, and zinc oxide) on alumina (Hardcastle and Wachs,
unpublished results) demonstrating that these metal oxides are also absorbed into the
surface of the alumina support.
6.2.3 Solid-state Chemistrv of SuDDorted Metal Oxides. The ability to discriminate
between the multiple states present in supported metal oxides with Raman spectroscopy
provides a convenient probe to monitor the structural transformations that occur in such
systems with increasing temperature. The transformations that occur for 10% WO,
supported on Al,O, as the calcination temperature is increased is shown in the Raman
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 135
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

H HH H
\o/ \/
I I
I
I I
I -H20
Ni+2 Ni+2 Ni+2 Ni+2 Ni+2 Ni+2
Downloaded by McGill University on 26 December 2012

Figure 14 Schematic of surface nickel oxide in the yAI,O, support illustrating its
inaccessibilityto surface moisture.

spectra of Figure 15.70 The 10%W03/A1203 sample treated at 500 OC exhibits only the
Raman bands of the surface tungsten-oxide phase. After heating to 950 "C additional
Raman bands appear at 253, 559, 743, and 840 cm-" due to the formation of @-A1203
which, unlike y-A1203,gives rise to weak Raman signals. Furthermore, the major Raman
band of the surface tungsten-oxide shifts from 986 to lo00 cm-* during this treatment as

Figure 15 Raman spectra of 10% WOJAi,O, as a function of calcination temperature.


View Online
136 Catalysis

a result of the increase in surface density of the tungsten-oxide overlayer on the alumina
support surface due to the significant decrease in surface area (from 180 to 65 m2/g) of
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

the alumina support. The surface tungsten-oxide Raman band at lo00 cm-' corresponds
to that found for exactly a monolayer of surface tungsten oxide on alumina?' These
structural transformations are schematically shown in Figures 16(a) and 16(b). Further
heating of the 10% W03/A1203sample to lo00 "C results in additional loss in surface
area of the alumina support, and a portion of the surface tungsten-oxide phase is forced
to leave the monolayer to form crystalline WO, (major Raman bands at 808,711, and 273
cm-') and crystalline A12(W04), (major Raman bands at 1057,386, and 371 cm-'). The
Downloaded by McGill University on 26 December 2012

spectrum of 10% W0,/A1203 heated to lo00 "C is dominated by the Raman bands of
crystalline WO, because of the extremely high Raman cross section of this oxide phase,
as discussed earlier. After additional heating at 1050 "C, all of the tungsten-oxide phases
in the 10% W03/A1203 transforms to A12(W04), -- the only thermally stable tungsten-
oxide compound for the W-Al-0 system. These changes are presented schematically in
Figures 16(c) and 16(d).

Flgure 16 Structural transformations of 10% WOdAI 0 system wlth calclnatlon


temperature. Structure after heatlng to: (a) 500 "C, (bj 850 "C, (c) 1000 "C, and (d)
1050 "C.
More detailed studies7' demonstrate that crystalline A12(W04)3is produced from
the direct solid-statereaction between crystalline WO, and the alumina support and occurs
by the migration of alumina to the WO, particles at these elevated temperatures. The
initial transformation of the surface tungsten-oxide phase to crystalline WO, particles at
elevated temperatures, when the alumina surface area decreases, reveals that the tungsten-
oxide species are not miscible in the alumina support and are actually diffusing away from
the alumina support. Thus, it appears that the surface tungsten oxide-aluminamixed oxide
forms an "immiscible oxide system" because of the inability of the alumina support to
accommodate the large charge and size of the @ cation. More recent Raman studies
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 137

with other alumina-supported surface oxides (Re7', Cr&, Mo&, V5+,Nb5', Ta5+, Ti&,
etc.) demonstrate that all surface metal oxides possessing large charged cations form
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

immiscible oxide systems with the alumina ~ u p p o r t . ' ~ ~ ' ~ * ~ ~


Similar Raman studies with small charged cations (Co2+,Ni2+, Cu2+, and Zn2+)
reveal that these cations form "miscible oxide systems" with the alumina support at
elevated temperatures (Hardcastle and Wachs, unpublished results). The transformations
that occur for 10%NiO supported on y-A1203as the calcination temperature is increased
are represented schematically in Figure 17. The in situ Raman studies have demonstrated
that, unlike the surface tungsten oxide-alumina system, the surface nickel-oxide species
Downloaded by McGill University on 26 December 2012

penetrate the alumina support surface and are incorporated within the surface layer after
calcination at 250 to 500 "C (see Figure 17(a)). Additional calcination at 700 "C provides
the necessary thermal mobility for the diffusion of the Ni2+cations from the surface layer
to the bulk region of the alumina support to form a solid solution of NiO-Al,O, (see
Figure 17(b)). Further calcination at temperatures in excess of 900 "C results in the
simultaneous crystallization of NiA120, and a-A1203 phases from the NiO-AI2O3 solid
solution (see Figure 17(c)). The formation of a-A1203at -900 "C is rather surprising
because crystalline a-A1203 usually forms at temperatures exceeding 1200 "C and
suggests that the a-A1203 phase may be produced by epitaxially growing on the NiAI20,
spinel particles. The formation of the a-A1203 phase at 900 "C is characteristic of the
miscible oxide systems (Co2+,Ni2+, Cu2+,and Zn2+). In contrast, the immiscible oxide
systems (Re7+,Cr&, Mo&, p, V5+, Nb5', Ta5', Ti&, etc.) form the 0-A120, phase at
these temperatures. These Raman studies reveal that the phase transformations of the
alumina support are determined by the nature of the surface metal oxide-support
interactions.

NV2
yA1203 Ni+2
NV2 Ni+'

Figure 17 Structural transformation of nlckel oxide/aiumlna as a function of calcination


temperature.
View Online
138 Catalysis

6.2.4 Molecular Structures of Surface Metal Oxides. Significantprogress has taken place
in recent years in detehning the molecular structures of surface metal oxides by Raman
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

spectroscopy. The fmt significant advance was the realization that under ambient
conditions the surface metal oxides are hydrated by adsorbed moisture and that the
molecular structures are different under ambient and in situ dehydrated condition^!^*^'
More recent studies have shown that the surface metal-oxide molecular structures present
under ambient conditions are the same metal-oxide species that are present in aqueous
solutions and that the solution pH controls the specific molecular ~tructures?~The pH
of the hydrated surface metal-oxide phase is determined by the specific oxide support and
Downloaded by McGill University on 26 December 2012

the surface coverage of the surface metal oxide and is directly related to the surface pH
at point zero charge (the pH at which there is a net surface zero charge). Thus, supports
such as magnesium oxide that possess a surface pH of about 12 at point zero charge form
solvated metal-oxide species typically observed in basic solutions. A hydrated alumina
support has a surface pH of about 8 at point zero surface charge and is slightly basic; a
hydrated titania support has a surface pH of about 6 at point zero surface charge and is
essentially neutral; and a hydrated silica support that possesses a surface pH of about 3
at point zero surface charge is somewhat acidic. As an example of how solution pH
affects the structure of a surface metal oxide species, the Raman spectra of chromium-
oxide reference compounds and supported chromium-oxide catalysts are presented below.
In aqueous solution chromium oxide can be present as CrO:-, Cr20,2-, Cr3O1;-,
and Cr401t-,and their relative concentrations depend on the solution pH and chromium
oxide concentrations. Monomeric species are dominant in basic and dilute solutions, and
polymerized species are dominant in acidic and concentrated solutions. The Raman
spectra of the chromium oxide ions are illustrated schematically in Figure 18 and their
vibrational assignments are presented in Table 6.88*89In general, chromium (VI) oxide
compounds possess strong Raman bands in the 800 to lo00 cm-' region due to Cr-0
stretching modes and bands of moderate intensity in the 300 to 400 cm-' region due to the
Cr-0 bending modes. These Raman band positions are consistent with tetrahedral
coordination for the aqueous chromium oxide ions. Tetrahedrally coordinated molecules
of the type CrOz' possess four Raman-active fundamental modes of vibration: symmetric
(Al) and antisymmetric (F2)stretching modes, and two bending modes (E and F2). In the
case of tetrahedral dimers, trimers and tetramers (see Figure 18 and Table 6), bands appear
in the Raman spectra that are due to Cr-0-Cr linkages (772-844 cm", 525-557 cm", and
209-217 cm-') and CrO, chain moieties (956-987 cm-'), both resulting from polymeriza-
tion of the monomeric chromate ~ n i t s . 8Similar
~ Raman spectra are observed for solid
chromium (VI)oxide compounds possessing monomers, dimers, trimers, and tetramers,
but the Raman spectra for the solid compounds contain additional internal modes due to
the slight distortion of the chromium oxide units in the solid state, as well as external
modes of the crystalline lattice.% The Raman spectra of crystalline Cr203 and
Cro~~All~7s03, which contain C?' cations, are presented in Figure 19. The Cro.25All.7s03
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 139
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

904 - '0
0
' 31 1 348
I I I
Downloaded by McGill University on 26 December 2012

348
987
I 525 I 318 'On

Flgure 18 Raman spectra of aqueous chromlum(V1) oxldes.88

Table 6 Raman band assignments for chromium (VI)oxide compounds


Band assignments' Monomer Dimer Trimer Tetramer
987 987
956 963
904 942 942b 942b
846 904 904 902
772 (vw)' 844 842
557 524 525
37 1 367 37 8 365
348 366 348
318
217 214 209
"as = antisymmetric stretch, ss = symmetric stretch, b = bending
h i m a t e d Raman stretching frequencies
Cvw = very weak

reference material contains C3' in solid solution with a-Al,O, (corundum). The C2'
reference compounds exhibit Raman bands in the 500 to 600 cm-' region due to their
lower Cr-0 bond strengths (resulting from the lower formal oxidation state) relative to the
Cr& compounds that exhibit Raman bands in the 800 to loo0 cm-' region. Thus, the
View Online
140 Catalysis
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

1200 1000 800 600 400 200


Raman Shift (cm-l)
Flgure 19 Raman spectra of solid chromium(Vi) oxides.

above discussion shows that Raman spectroscopy can discriminate between various
chromium-oxide compounds as well as provide direct structural information about each
compound.
The interaction of chromium oxide with different oxide supports (A1203,TiO,, and
SiO,) under ambient conditions, where the surface metal oxide phases are hydrated, is
reflected in the Raman spectra presented in Figure 20. The nature of the supported
chromium-oxide phases is determined from a comparison of the Raman spectra of the
supported chromium-oxide phases and the reference chromium-oxide compounds already
discussed. The 4% Cr03/A1,03 Raman spectra possess Raman features at 890 cm-' -
symmetric stretch, 930 cm" - antisymmetric stretch, and 363 cm-' - bending, which are
consistent with an isolated, tetrahedral chromate species [Cr04]. In addition, there is a
very weak band at 217 cm-' due to the bending mode of the Cr-O-Cr linkage which is
diagnostic for the presence of a small amount of dichromate species [Cr207]. The 4%
Cr03/Ti02 Raman spectrum possesses the same features as the 4% Cr03/A1203sample
in the 800 to loo0 cm" region and suggests that chromium oxide on titania is also present
as an isolated, tetrahedral chromate species [CrO,J. The inability to obtain chromium
oxide Raman data in the 200 to 300 cm-' region, because of the strong TiO, Raman
bands, prevents the direct identification of any dichromate species. The similar Raman
features in the 800 to 900 cm-' region for the 4% Cr03/A1203 and 4% Cr03/Ti02,
however, strongly suggest that the chromium oxide species are similar in both of these
systems.
The Raman spectrum for the 4% CrO,/SiO, sample is very different from that
obtained for 4% Cr03/A1203and 4% Cr03/Ti02 and reflects the presence of a more
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 141
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

1
I I I I I I 1 1 -

1200 1000 DO0 600 400 200

Raman Shift (cm-l)

Figure 20 Raman spectra of 4% chromium oxide supported on alumina, titanla, and silica.

complex supported chromium-oxide phase on SiO,. The 4% CrO,/SiO, Raman bands


(963 cm-'-symmetric stretch of the nonterminalCr0, unit, 902 cm"-symmetric stretch of
the terminal CrO, unit, 844 cm-'-antisymmetric stretch of the Cr-0-Cr linkages (or Cr-0
stretch of CrOH functionalities), 370 cm-'- bending mode of the CrO, unit, and 217 cm"-
bending modes of the Cr-0-Cr linkage are consistent with the presence of tetrachromate
s p i e s [Cr4013]and Cr203crystallites(diagnostic band at 543 cm-'). The Raman feature
at 989 cm-' is due to the formation of a strongly interacting chromate species resulting
from the partial dehydration of the surface by the laser beam.', The ambient Raman
studies of chromium oxide on Al,03,TiO,, and SiO, supports reveal that the oxide
support has a strong effect on the nature of the supported chromium-oxide phase which
is related to the point zero surface charge of the hydrated oxide support (see reference 13
for a more detailed investigation of supported chromium-oxidecatalysts). Similar ambient
Raman studies on the influence of oxide supports upon supported rhenium oxide, tungsten
oxide, molybdenum oxide, vanadium oxide, and niobium oxide are presented in references
71, 82, and 91.
Under dehydrated conditions, the adsorbed moisture is removed and the in situ
Raman spectra of the surface metal oxides differ markedly showing that the structures of
the dehydrated species are very different from those of their hydrated counterparts (see
references above in Section 6.2.2). These changes are not surprising since the influence
of the net zero surface charge of the oxide support can only be exerted in an aqueous
medium. For the dehydrated surface metal oxides, however, essentially the same
molecular structures are seen on a l l the oxide supports for each supported metal oxide.92
View Online
142 Catalysis

The structures of the dehydrated surface metal oxides are highly distorted and the absence
of appropriate reference compounds sometimes complicates the structural assignment.
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

However, with the aid of complementary spectroscopies it has been possible to assign the
structures of the dehydrated surface metal-oxide specie^.'^-'^^^^^ In general, two different
types of dehydrated surface metal-oxide species have been identified: a highly distorted
species, which is relatively more abundant at lower coverages, and a moderately distorted
species, which becomes more abundant at higher coverages. The highly distorted species
typically terminates the surface with one very strong metal-oxygen bond, M=O (two
terminal bonds for chromates), and is regarded as an isolated surface species. The slightly
Downloaded by McGill University on 26 December 2012

distorted surface species, in comparison, has metal-oxygen bonds of lower bond order,
suggesting a non-terminating functionality,and is present as a polymerized species. More
research will be required before a model to predict the dehydrated surface metal-oxide
structures can be established.

6.3 Zeolites. - Extensive Raman investigations of aluminosilicate zeolitic materials have


been initiated only in the last few years. Raman spectroscopy of zeolites offers certain
advantages over more conventional vibrational methods. For example, the extremely low
Raman scattering from water molecules allows for the direct monitoring of the events that
occur in solution during zeolite synthesis. The weak Raman signals of aluminosilicate
materials result in a weak substrate background spectrum and allow for the observation
of the vibrational spectra of adsorbed molecules as well as inorganic ions present in the
zeolite structure below 1200 cm-' (discussed below in Section 6.4). Furthermore, the
weak Raman vibrations of the zeolite framework can also be measured and provide
fundamental information about the aluminosilicate structure.
6.3.1 Zeolite Framework. Experimental problems with zeolite sample fluorescence
initially dampened the widespread application of Raman spectroscopy to ze0lites.9~ In
spite of these experimental complications, many Raman papers on zeolite structure have
begun to appear in the literature.w-'06 The fluorescence is thought to originate from
transition metal cations (Fe3', C?', and Ah2') and organic impurities. Recent advances
in Raman instrumentation, such as FT-Raman spectroscopy, time-resolved techniques,
resonance Raman spectroscopy,and CCD detectors (allowing red/near infrared excitation)
will hopefully reduce the effect of fluorescence on the Raman spectra of zeolites and
should contribute to greater interest in Raman spectroscopy of zeolitic materials.
The influence of exchangeable cations Li', Na', , 'K , '
l
T and NH4' on the
vibrations of hydrated zeolite A (composed of alternating, comer-sharing SiO, and AlO,
tetrahedra) was investigated by Raman spectro~copy?~The Raman spectra are presented
in Figure Spectra below 300 cm" were not collected because of considerable light
scattering by the zeolite crystals. The Raman bands in the 800 to 1100 cm-' region were
assigned to the stretching vibrations of the SiO, groups, the bands in the 600 to 800 cm-'
region were assigned to the stretching vibrations of the A10, groups, and the bands in the
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 143
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

Repmdwd with permission of


the American Chemical Society.

Raman Shift (cm-')


Flgure 21 Raman spectra of hydrated zeollte A exchanged with LP, Na', K+, TI+, and NH,+
Ions: laser Ilne, 457.9 nrn; laser power, 150 mW; sllt wldth, 5 cm". The dots denote
plasma liness

300 to 700 cm-' region were assigned to the bending vibrations of the SiO, groups. The
weaker bending vibrations of the A104 groups are not observed because of the ionicity of
the Al-0 bonds and somewhat smaller bond order present in the AlO, groups relative to
the SiO, groups. Information about the vibrations of the oxygen linkages in Al-O-Si that
would be expected to occur below 300 cm-' was not obtained, as mentioned above.
The Raman data reveal that there are some significant changes in the zeolite A
vibrational spectrum as Na' is replaced by Li', K+,Tl+ and N H,'. The similar Raman
spectra of Na', K+,and T1+ containing zeolites suggest that no major distortion of the
lattice takes place and that the cations occupy similar sites. In contrast, the Li+ cation
tends to distort the zeolite A framework structure and cause a decrease in the Si-O-A1
bond angles relative to Na+-zeolite A. The NH,+-zeolite A exhibits an increase in
frequency of the vibrations which indicates hydrogen bonding between the cation and the
zeolite oxygens. These studies reveal the fundamental structural information obtained
from Raman examination of zeolitic materials.
The dependence of the framework vibrational frequencies on the Si/Al ratio and
aluminosilicate ring size were also examined with Raman spectroscopy, shown in
Figure 22.'O0*'*' The Si/Al ratio was varied from 1.0 to 2.7 in a series of zeolite A
materials. The low-frequency bands at 337 and 410 cm-' were not found to change with
the Si/Al ratio. The strong band at 489 cm-' exhibits a weak dependence on Si/Al ratio.
The 700 cm-' band, however, shows the most rapid and almost linear increase in
frequency with Si/Al ratios. The bands in the 900 to 1100 cm-' region exhibit a
complicated dependence on the Si/Al ratio. The strong Raman band at about 500 cm-',
which possesses a weak dependence on the Si/Al ratio, however, is very sensitive to the
View Online
144 Catalysis
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

Reproduced with permission of


the American Chemical Society.
Raman Shift (cm-I)
Flgure 22 Raman spectra of calcined and Na+-exchangedzeollte samples: (a) SI/AI = 1.O;
(b) SUAl = 1.2; (c) SI/AI = 1.4; (d) SI/AI = 2.0; (e) SVAI t 2.7. Laser line = 457.9 nm,’w

Si-0-Al angle in the zeolite structure. It appears that as the Si-O-A1 angle increases, the
frequency of the bending mode decreases for zeolites built up of four-membered rings.
Thus, the Raman spectrum is sensitive to the arrangement of the Si and A1 atoms and the
coupling between them.
6.3.2 Svnthesis of Zeolites. A better molecular level understanding of the zeolite
synthesis steps occurring in solution should assist in the development of new zeolitic
materials. Raman spectroscopy offers the advantage that it can provide fundamental
information about the aqueous phase as well as the solid amorphous and crystalline phases
present during the zeolite synthesis. Several Raman studies on the synthesis of
aluminosilicate zeolitic materials have recently been rep~rted.”*~’-~
The transformation of aluminosilicate gel to zeolite A was investigated by Raman
spectro~copy.~~ The amorphous gel has a structure consisting of predominantly four-
membered rings connected in a random fashion. It is considerably depolymerized,
consisting of Si atoms with one or two nonbonded oxygen atoms. The transformation of
the gel to zeolite A proceeds only in the presence of Al(OH),- species in the solution.
The gel reorganizes its structure by interaction with the Al(OH),- ions and forms nuclei
of zeolite A. The crystallization curve obtained by Raman spectroscopy closely resembles
that from X-ray diffraction. Analogous investigations examining the synthesis of zeolites
X, Y,and Z-SM5 have also appeared.”. 98* 99

6.4 Chemisomtion Studies. - Infrared spectroscopy is usually the technique of choice in


chemisorption studies on catalytic surfaces because of its rapid acquisition of data, good
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 145

sensitivity, and nonintrusive radiant flux. Recent Raman investigations, however, have
shown that Raman spectroscopy offers certain advantages that are not achievable with
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

infrared spectroscopy. Typical catalytic supports such as Si02 and A1203possess very
strong infrared absorptions below 1200 cm-' and the relatively weak Raman scattering
from such catalyst supports facilitates detection of the vibrational spectra of chemisorbed
species below 1200 cm-'. Furthermore, the far infrared region, <400 cm-', is easily
measured with Raman spectroscopy. The adsorbate/solid interface in aqueous media can
also be monitored with Raman spectroscopy because, unlike infrared absorption, the
Raman signal is not significantly affected by aqueous solvents. The different selection
Downloaded by McGill University on 26 December 2012

rules for Raman and infrared spectroscopies allow for a greater number of the vibrational
modes to be measured and consequently, a more complete structural assignment of the
adsorbed species may be made by utilizing the complementary information provided by
Raman and infrared spectroscopies. The application of Raman spectroscopy to
chemisorption studies will only be briefly discussed here since an extensive review of this
subject appeared in an article by Bartlett and C ~ o n e y . ~
e d Catalvsts. In supported metal catalysts, the oxide supports as well
6.4.1 S u ~ ~ o r t Metal
as the supported metal components usually do not give rise to Raman vibrations. Such
Raman studies, however, have typically been characterized by low signal-to-noise ratios.
The recent advances in Raman instrumentation should significantly improve the quality
of data obtained in such chemisorption studies. Silica-supported nickel catalysts have
been examined for chemisorption of CO, H2, C2H4,and ben~ene.''~-''~ When ethylene
was adsorbed, vibrational bands assigned to v(C-H), v(C-C), G(C-CH), G(Ni-H), and v(Ni-
C) were detected. The analysis showed that ethylene is adsorbed associatively at low
temperatures and dissociatively at room temperature where trimerization of adsorbed
species to benzene takes place."' Adsorption of H, gave rise to three bands which were
assigned to v(Ni-H), G(Ni-H), and v(Ni-H) for multibonded Chemisorption
of CO exhibited vibrations due to v(C-0) from linear and bridged sites as well as bands
due to v(Ni-C) and G(Ni-C=O) mode^.''^^"' These few studies reveal the comprehensive
vibrational information provided by Raman spectroscopy in chemisorption studies on
supported metal catalysts.
6.4.2 Bulk Metal Oxides. Extensive Raman chemisorption studies on high surface area
alumina and silica supports have been performed because of the industrial importance of
these oxides and their weak background Raman vibration^.^ The most informative studies
resulted from the adsorption of pyridine since this probe molecule is very sensitive to the
type of acid sites (Bransted and Lewis) present on the silica and alumina surfaces. On
the alumia support, Lewis pyridine was predominately ~bserved''~and on the silica
support both Lewis and Brgnsted pyridine were ob~erved."~The surface concentrations
of pyridine on the silica surface were very small in comparison to the pyridine coverages
on the alumina surface. The signal intensities were dramatically enhanced by the
View Online
146 Catalysis

application of resonance Raman spectroscopy, and it was estimated that only lo-, to
of a pyridine monolayer was present on the silica surface.
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

Raman chemisorption studies on TiO,, ZnO, and SnO, supports have also been
reported? The low surface areas of these oxide supports and their strong Raman
scattering (below 700 cm-') have led to the use of resonance Raman spectroscopy in
chemisorption investigations on these oxides. The time frame of surface reactions
occurring on colloidal TiO, surfaces during photochemical processes is to lo-',
seconds, and time-resolved resonance Raman spectroscopy has been used to elucidate the
reaction kinetics and mechanisms in such systems."5-' l7 Thus, Raman spectroscopy
Downloaded by McGill University on 26 December 2012

provides detailed insights into the nature of surface processes that would be extremely
difficult to study by other techniques.
6.4.3 S u ~ ~ o r t Metal
e d Oxide Catalvsts. Raman chemisorption studies on supported metal
oxides, utilizing different probe molecules (i.e., pyridine, thiophene, HCl, HBr, etc.) have
been undertaken in order to learn about the nature of the metal-oxide support interactions
in such systems. A series of detailed studies on the adsorption of pyridine on
MoO,/Al,O, catalysts were reported!2-' ''-'l9 These studies concluded that Brgnsted
pyridine was associated with the surface molybdenum-oxide species, and that the
molybdena-alumina interaction occurs preferentially at bridging hydroxyl groups on
octahedrally coordinated aluminum atoms since Lewis Raman bands arising from pyridine
adsorbed at such sites decrease in intensity with increasing molybdenum oxide coverage.
The formation of significant amounts of crystalline MOO, was found to inhibit the
adsorption of pyridine by blocking both Lewis and Brgnsted acid sites. Addition of
cobalt-oxide promoter eliminated the Lewis pyridine Raman band suggesting that Co2+
ions interact with the aluminum sites responsible for Lewis acidity.
6.4.4 Zeolites. The weak Raman signals arising from the aluminosilicate zeolite
framework allow for the detection of vibrational bands of adsorbates, especially below
1200 cm", which are not readily accessible to infrared absorption techniques. Raman
spectroscopy is an extremely effective characterization method when two or more colored
species coexist on the surface, since the spectrum of one of the species may be enhanced
selectively by a careful choice of the exciting line.'20 A wide range of adsorbate/zeolite
systems have been examined by Raman spectroscopy and include SO,, NO,, acety-
lene/polyacetylene, dimethylacetylene, benzene, pyridine, pyrazine, cyclopropane, and
halogens. Extensive discussions of these absorbate/zeolite studies are found in a review
article by Bartlett and Cooney?
The state of the template tetrapropylammonium ion [ P A ] ' was also monitored
with Raman spectroscopy during various stages of ZSM-5 ~ y n t h e s i s . ~ ~ *It' was
~ ' ~found
'~~
that this cation is trapped into the amorphous solid phase at the earliest stages of the
synthesis in the all-trans configuration. Upon crystallization of the zeolite, there is a
forced change in the conformation of the trapped tetrapropylammonium cation such that
it can fit into the zig-zag zeolite channels. The template [PA]' was decomposed by
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 147

sample calcination at temperatures in excess of 673 K. Thus, Raman spectroscopy


provides more fundamental information about the nature of adsorbate/zeolite interactions
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

than other vibrational techniques.


6.4.5 Surface Science Studies. The recent advances in state-of-the-art Raman instrumen-
tation, such as triple monochromators with high stray-light rejection ratios, intensified and
unintensified detectors with multichannel capability (IPDAs and CCDs), have led to the
emergence of Raman spectroscopy as a surface science tool (i.e., the detection and
quantification of very low concentrations of molecules on low surface area single-crystal
Downloaded by McGill University on 26 December 2012

surfaces). Submonolayer sensitivity of adsorbates on single crystals is essentially


impossible without multichannel d e t e ~ t i 0 n . l ~For
~ example, the Raman signal can be
estimated from
SRmm = n L (ik/8st)st
for a monolayer of benzene on a single-crystal where CT is the Raman scattering cross
section, 52 is the solid angle subtended by the collection optics, (ik/6s2) is the differential
scattering cross section, which is approximately cm2 steradians-' molecule-' (47c
steradians is the solid angle subtended by a sphere), n is the number of scatterers in the
probed region (about 10" benzene molecules per cm2), L is the laser flux (3 X 102'
photons/sec), and R is the collection angle (which is about 1 steradian). This results in
about 300 photons per second being collected by the optics which is directed into a
spectrometer. If the spectrometer is roughly 10% efficient and the detector possesses a
quantum efficiency of about lo%, then only three photons are detected per second. If the
desired signal-to-noiseratio is 10, for a 3000 cm-' scan with a resolution of about 5 cm-',
the total time of scanning takes about 5 hours. This time scale is generally unacceptable
for ultra-high vacuum experiments where sample integrity degrades with time. In
comparison, only 30 seconds is required for a 600channel OMA for the same resolution,
and the savings in time is a factor of 600. Therefore, OMA detection offers a tremendous
experimental advantage over the conventional scanning spectrometer for the detection of
submonolayer amounts of absorbed species on single crystal^.'^^*'^^
The design considerations for a surface Raman spectrometer for normal
(unenhanced) Raman spectroscopy adsorbed on surfaces have been thoroughly reviewed
by Campi~n.'~' The Raman instrumental design was based on the behavior of the
electromagnetic field near the conducting surface of the single crystal, and the detection
was carried out with an optical multichannel detector to significantly reduce the time
required to accumulate the spectrum. The angle of observation is critical in these
experiments since the orientation of the induced dipole (of an adsorbed molecule), as well
as the polarization of the incident field, determines the variation in scattered intensity with
observation angle.
The Raman spectrum of a thin film (-5 nm) of nitrobenzene physically adsorbed
on a Ni( 111) surface is shown in Figure 23.'26 This spectrum would take on the order
of a few days to collect with a scanning instrument but was acquired in less than 10
View Online
148 Catalysis
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102
Downloaded by McGill University on 26 December 2012

Reproduced with permission of I . I . t . I . l . I . I 0 1

the American Chemical Society. 2zw 2000 1000 1600 14w IZW low BW

AV (cm

Flgure 23 Normal Raman spectrum of a very thin film (-5 nm) of nitrobenzenecondensed
on a Ni (111) surface at 100 K. Normal Raman spectrum of less than a monolayer of
nltrosobenrene adsorbed on Nl(ll1) at 100 K.'=

minutes using about 150 mW of laser power and OMA detection. The observed
differences between the spectra suggest that some chemical changes have occurred to the
nitrobenzene molecule upon adsorption.
The Raman spectrum of ethylene chemisorbed on Ag( 110) is shown in Figure 24
and provides an example of catalytic interest.'26 The C=C stretch of adsorbed ethylene
is observed at 1580 cm-' and the symmetric C-H in-plane deformation at 1320 cm".

1800 1700 1000 1500 1400 1300 1200 1100

AV (cm -l)
Reproduced with permission of the American Chemical Society.
Flgure 24 The normal Raman spectrum of ethylene chemlsorbedon Ag (110) at submono-
layer coverage at 100 K. The silver surface was predosed with 0, at 300 K>=
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 149

These observed frequencies are about 20 cm-' lower than for ethylene in the gas phase and
suggest a weak x-bonding interaction with the Ag surface. Further experiments show that
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

the ethylene adsorbs on a silver atom and not on an oxygen atom (which was preadsorbed
onto the
6.4.6 Surface-enhanced Raman Scattering. Surface-enhanced Raman scattering (SERS)
has recently gained much attention, including its use in the field of heterogeneous
catalysis. SERS was fmt observed by Fleischman et al.12' who found enhanced signals
from pyridine adsorbed on an electrochemically roughened silver electrode surface but
Downloaded by McGill University on 26 December 2012

attributed the enhancement to the increased surface area caused by the roughening. A true
enhancement mechanism was suggested later.'28 The mechanism behind the enhancement
has been the subject of many Raman investigations carried out particularly on silver, gold,
and copper metals where the surface enhancement is large.
The extraordinary enhancement of the Raman spectrum of pyridine adsorbed on
silver has resulted in the investigation of pyridine as a candidate for possible enhancement
on other metals. More recently, SERS has been used to probe the adsorption of molecules
on metals that are of catalytic importance. For example, SERS spectra have been
collected of pyridine adsorbed on metallic rhodium in the fmt demonstration of surface-
~ ~ the enhancement of pyridine on silver is
enhancement from metallic r h 0 d i ~ m . IWhile
thought to arise from an electrodynamic enhancement, the enhancement of the pyridine
Raman spectrum by rhodium is thought to arise from a chemical effect since rhodium,
unlike silver, is not a free electron metal.
SERS has also been shown to be a useful analytical tool in the elucidation of
reaction mechanisms in heterogeneous catalysis (e.g., dehydroamination reactions on
copper surfaces).'30 The adsorption and desorption of amines were investigated on copper
foils which were roughened by chemical etching and exhibited strong surface enhance-
ment. Enhanced Raman spectra were obtained for adsorbed m-toluidine on copper with
submonolayer sensitivity. The mechanism of adsorption is thought to be via the nitrogen
lone pair without N-Hbond dissociation.

7 Future of Raman Spectroscopv

The future of Raman spectroscopy in the research and the development of catalysts
appears to be extremely promising. The recent revolution in Raman instrumentation has
dramatically increased the ability to detect weak Raman signals and to collect the data in
very short times. Thus, it is now possible to perform real-time Raman analysis and to
study many catalytic systems that give rise to unusually weak Raman signals. The
enormous strides in Raman instrumentation now allow for the characterization of a wide
range of catalytic materials: bulk mixed oxides, supported metal oxides, zeolites, supported
metal systems, metal foils, as well as single crystal surfaces. Few Raman studies have
been reported for sulfides, nitrides, or carbides, but these catalytic materials also give rise
View Online
150 Cutalysis

to Raman signals. The potential of modern Raman spectroscopy in heterogeneous


catalysis is only beginning to be explored by catalytic scientists. It will probably require
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

another decade to realize the full potential of the current state-of-the-art Raman
instrumentation. Continuing advances in the technology of Raman spectroscopy are
expected to open up new research opportunities that are not accessible with the current
generation of Raman instrumentation.
The strength of Raman spectroscopy originates in its ability to address many
fundamental and practical issues in heterogeneous catalysis that are extremely difficult or
even impossible to study with other catalyst characterization methods. For example, the
Downloaded by McGill University on 26 December 2012

synthesis of various catalytic materials can be directly investigated with Raman because
of the absence of or weak Raman signals from the solvents. The weak Raman signals
from many oxide-support materials (A1,0,, SiO,, etc.) allow for the collection of
vibrational information below 1200 cm-'; this information is solely responsible for the
recent advances in our understanding of supported metal-oxide catalysts. Similarly, the
ability to collect vibrational information of adsorbed molecules below 1200 cm-' provides
additional insights into the chemisorption of molecules on practical catalytic surfaces. The
molecular structural information provided by Raman spectroscopy is currently allowing
us to establish molecular structure-reactivity relationships and will assist us in the
molecular design of complex catalytic materials.

References

1 I.E. Wachs, F.D. Hardcastle, and S.S. Chan, Spectroscopy, 1986, 1, 30.
2 L. Dixit, D.L.Gerrard, and H.J. Bowley, Appl. Spectrosc. Rev., 1986, 22, 189.
3 J R. Bartlett and R.P. Cooney, in 'Spectroscopy of Inorganic-based Materials', ed. R.J.H.
Clark and R.E. Hester, John Wiley & Sons Ltd., 1987, p. 187.
4 J.M. Stencel, 'Raman Spectroscopy for Catalysis', Van Nostrand Reinhold, New York,
1990.
5 C.V. Raman and K.S. Krishnan, Nature, 1928, 121, 501.
6 E.B. Wilson, J.C. Decius, and P.C. Cross, 'Molecular Vibrations: The Theory of Infrared
and Raman Vibrational Spectra', Dover Publications, New York, 1980.
7 D.A. Long,'Raman Spectroscopy', McGraw-Hill; New York, 1977.
8 L. A. Woodward, in 'Raman Spectroscopy', Vol. 1, ed H.A. Szymanski, Plenum Press,
New York, 1967.
9 G. Placzek, M a n Handbuch der Radiologie, Vol. 6, Part 2, Akademische Verlagsge-
sellschaft M.B.H., Leipzig, 1934.
10 F.A. Cotton, 'Chemical Applications of Group Theory', Wiley-Interscience, New York,
1971.
11 K. Nakamoto, 'Infrared and Raman Spectra of Inorganic and Coordination Compounds',
John Wiley & Sons, New York, 1978.
12 F.D. Hardcastle, I.E. Wachs, J.A. Horsley, and G.H. Via, J. Mol. Cutal, 1988.46, 15.
13 F.D. Hardcastle and I.E. Wachs, J. Mol. Cutal, 1988, 46, 171.
14 F.D. Hardcastle and I.E. Wachs, J. Raman Spectrosc., 1990,21, 683.
15 F.D. Hardcastle and I.E. Wachs, J. Phys. Chem., 1991.95, 5031.
16 F.D. Hardcastle and I.E. Wachs, Solid State Ionics, 1991, 45, 201.
17 F.D. Hardcastle and I.E. Wachs, J. Solid State Chem.,1992,97, 319.
18 F.D. Hardcastle and I.E. Wachs, in preparation.
19 F.D. Hardcastle, C.H.F. Peden, and K.B. Kidd, J. Phys. Chem., 1992, submitted.
20 I.D. Brown, Chem. SOC.Rev., 1978, 7 , 359.
21 L. Pauling, J . Am. Chem. Suc., 1929, 51, 1010.
View Online
Applications of Rarnan Spectroscopy to Heterogeneous Catalysis 151

22 I.D. Brown and K.K. Wu, Actu Cryst., 1976,B32, 1957.


23 A.F. Corsmit, H.E. Hoefdraad, and G. Blasse, J. Inorg. Nucl. Chem., 1972,34,340
24 G. Blasse and A.F. Corsmit, J. Solid State Chem., 1973,6,513.
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

25 N.E. Schlotter, S.A. Schaertel, S.P. Kelty, and R. Howard, Appl. Spectrosc., 1988,
746.
26 A. Campion and S.S. Perry, Laser Focus World, 1990,26,113.
27 P.P. Yaney, J. Opt. SOC. Am., 1972,62,1297.
28 A.F. Van den Elzen and G.D. Rieck. Actu Cnst., 1973,B29, 2436.
29 J.M. Hanis, R.W. Chrisman, F.E. Lytle, and R.S: Tobias, Anal. Chem., 1976,48, 1937.
30 B. Chase and Y. Talmi, Appl. Spectrosc., 1991.45.929.
31 T. Hirschfeld and B. Chase, Appl. Spectrosc., 1986,40,133.
32 B. Chase, Anal. Chem., 1987,59, 811A.
33 E.N. Lewis, V.F. Kalasinsky, and I.W. Levin, Appl. Spectrosc., 1988,42, 1188.
Downloaded by McGill University on 26 December 2012

34 A. Crookell, M. Fleischrnann, M. Hanniet, and P.J. Hendra, Chem. Phys. Lett.,1988,149,


123.
35 B. Chase and B.A. Parkinson, Appl Spectrosc., 1988.42, 1186.
36 A.J. Sommer and J.E. Katon, Appl. Spectrosc., 1991,45,527.
37 D.C. Tilotta and W.G. Fateley, Spectroscopy, 1988,3,14.
38 D.C. Tilotta, R.D. Freeman, and W.G. Fateley, Appl. Spectrosc., 1987.41, 1280.
39 P.J. Treado, A. Govil, M.D. Moms, K.D. Stemitzke, and R.L. McCreery, Appl.
Spectrosc., 1990,44, 1270.
40 A. Govil, D.M. Pallister, L.-H. Chen, and M.D. Moms,Appl. Spectrosc., 1991,45,1604.
41 S.A. Asher, P. Flaugh, and G. Washinger, Spectroscopy, 1986,1,26.
42 C.P. Cheng, J.D. Ludowise, and G.L. Schrader, Appl. Spectrosc., 1980,34,146.
43 S.S. Chan and A.T. Bell, J. Cutul, 1984,89,433.
44 D.P. Strommen, and K. Nakamoto, 'Laboratory Raman Spectroscopy', John Wiley &
Sons, New York, 1984.
45 R.P. Durham and D.J. Wood, in 'Analytical Applications of Spectroscopy', eds. C.S.
Creaser and A.M.C. Davies, Royal Society of Chemistry, 1988.
46 R.P. Cooney, G. Curthoys, and T.T. Nguyen, Adv. Cutul., 1975,24, 293.
47 T.A. Egerton and A.H. Hardin, Catal. Rev. Sci.-Eng., 1975,11,71.
48 W.N. Delgass, G.L. Haller, R. Kellerman, and J.H. Lunsford, 'Spectroscopy in
Heterogeneous Catalysis', Academic Press, New York, 1979,p. 58.
49 B.A. Morrow, in 'Vibrational Spectroscopies for Adsorbed Species', ed A.T. Bell and
M.L. Hair, ACS Symposium Series 137, American Chemistry Society, Washington DC,
1980,p. 119.
50 J. Hauck and A. Fadini, 2. Nuturforsch, 1970,25b,422.
51 J.A. Horsley, LE. Wachs, J.M. Brown, G.H. Via and F.D. Hardcastle, J. Phys. Chem.,
1987,91,4014.
52 N. Weinstock, H. Schulze, and A. Muller, J. Chem. Phys., 1973,59, 5063.
53 H.J. Becher, H.J. Brockmeyer, and U . Prigge, J Chem. Res., 1978,M , 1670.
54 F.D. Hardcastle and I.E. Wachs, J. Phys. Chem., 1991,95, 10763.
55 R.K. Grasselli and J.D. Burrington, Adv. Cutul, 1981,30, 133.
56 R.K. Grasselli, J. Chem. Educ., 1986,63,216.
57 A.F. Van den Elzen and G.D. Rieck, Acta Crysr., 1973,B29, 2433.
58 H. Chen and A.W. Sleight, J. Solid State Chem.,1986,63, 70.
59 R.G. Teller, J.F. Brazdil, R.K. Grasselli, and J.D. Jorgensen, Actu Cryst., 1984, 40C,
2001.
60 D.J. Buttrey, D.A. Jefferson, and J.M. Thomas, Philos. Mug.A, 1986,53, 897.
61 D.J. Buttrey, D.A. Jefferson, and J.M. Thomas, Muter. Res. Bull., 1986,21, 739.
62 A. Watanabe, S. Horiuchi, and H. Kodarna, J. Solid State Chem., 1987,67,333.
63 S . Miyazawa, A. Kawana, H. Koizumi, and H. Iwasaki, Muter. Res. Bull., 1974,9,41.
64 G. Wandahl and A.N. Christensen, Actu Chem. Scund., 1987,41A,358.
65 M. Liegeois-Duyckaerts and P. Tarte, Spectrochim. Acta, 1972,28A, 2037.
66 J.F. Brazdil, M.Mehicic, L.C. Glaeser, M.A.S. Hazle, and R.K. Grasselli, in 'Catalyst
Characterization Science', eds. M.L.Deviney and J.L. Gland, ACS Syrnp. Series 288,
1985.
67 L.C. Glaeser, J.F. Brazdil, M.A. Hazle, M. Mehicic, and R.K. Grasselli, J. Chem. SOC.,
Furuduy Trans. I , 81,2903.
68 S.S.Chan, I.E. Wachs, and L.L. Murrell, J. Catul, 1984,90, 150.
View Online
152 Catalysis

69 S.S. Chan, I.E. Wachs, L.L.Murrell, L. Wang, and W.K. Hall, J. Phys. Chem., 1984,88,
5831.
70 S.S. Chan, I.E. Wachs, L.L. Murrell, and N.C. Dispenziere, J. Cafal, 1985, 92, 1.
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

71 G. Deo and I.E. Wachs, J. Phys. Chem., 1991,95, 5889.


72 K. Segawa and W.K. Hall, J. Cafal, 1982.77, 221.
73 J.M. Stencel, L.E. Makovsky, J.R. Diehl, and T.A. Sarkus, J. Raman Specfrosc., 1984.15,
282.
74 J.M. Stencel, L.E. Makovsky, J.R. Diehl, andT.A. Sarkus, J. Cafal, 1985, 95, 414.
75 E. Payen, S. Kasztelan, J. Grimblot, and J.P. Bonnefle, J. Raman Specfrosc., 1984, 15,
282.
76 E. Payen, S. Kasztelan, J. Grimblot, and J.P. Bonnelle, J. Raman Specfrosc., 1986, 17,
233.
77 L. Wang and W.K. Hall, J. Cafal, 1983, 82, 177.
Downloaded by McGill University on 26 December 2012

78 C. Cristiani, P. Fonatti, and G. Busca, J. Cafal, 1989,116, 586.


79 S.T. Oyama, G. Went, K.B. Lewis, A.T. Bell, andG. Somorjai, J. Phys. Chem., 1989,93,
6786.
80 I.E. Wachs, J. Cafal, 1990, 124, 570.
81 G.T. Went, T.S. Oyama, and A.T. Bell, J. Phys. Chem., 1990.94, 4240.
82 J.-M. Jehng and I.E. Wachs, Cafal. Today, 1990, 8, 37.
83 J.-M. Jehng and I.E. Wachs, J. Phys. Chem., 1991, 95, 7373.
84 T. Machej, J. Haber, A.M. Turek, and I.E. Wachs, Appl. Cafal., 1991, 70, 115.
85 M.A. Vuurman, I.E. Wachs, and A.M. Hirt, J. Phys. Chem., 1991,95, 9928.
86 C.H. Perry, E. Anastassakis, and J. Sokoloff, Ind. J. Pure. Appl. Phys., 1971, 9, 930.
87 M.A. Vuurman and I.E. Wachs, J. Phys. Chem., 1992, 96.
88 G. Michel and R. Cahay, J. Raman Specfrosc., 1986, 17.4.
89 G. Michel and R. Machiroux, J. Raman Specfrosc., 1983, 14, 22.
90 R. Mattes, 2.Anorg. Allg. Chem., 1971, 392, 163.
91 I.E. Wachs and F.D. Hardcastle, in 'Proc. 9th Intern. Congr., Cafal.', 1988, 4, 1449.
92 I.E. Wachs, G. Deo, M.A. Vuurman, D.S. Kim, and H. Hu, in 'Proc. 10th Intern. Congr.
Catal.', 1992, in press.
93 C.L. Angell, J. Phys. Chem., 1973, 77, 222.
94 F. Roozeboom, H.E. Robson, and S . S . Chan, Zeolites, 1983, 3, 321.
95 P.K. Dutta and B. Del Barco, J. Chem. SOC., Chem. Commun., 1985, 1297.
96 P.K. Dutta and B. Del Barco, J. Phys. Chem., 1985, 89, 1861.
97 P.K. Dutta and D.C. Shieh, J. Phys. Chem., 1986, 90, 2331.
98 P.K. Dutta, D.C. Shieh, andM. Puri, J. Phys. Chem., 1987,91, 2332.
99 P.K. Dutta and M. Puri, J. Phys. Chem., 1987,91,4329.
100 P.K. Dutta and B. Del Barco, J. Phys. Chem., 1988,92, 354.
101 P.K. Dutta and R.E. Zaykoski, Zeolites, 1988.8, 179.
102 P.K. Dutta, D.C. Shieh, and M. Puri, Zeolites, 1988, 8, 306.
103 P.K. Dutta and M. Puri, Proc. Mat. Res. SOC. Symp., 111, 101.
104 P.K. Dutta and M. Puri, J. Cafal, 1988, 111, 453.
105 P.K. Dutta, M. Puri, and C. Bowers, ACS Symp. Series, 1989.
106 P.K. Dutta and R.E. Zaykoski, J. Phys. Chem., 1989.
107 W. Krasser, A. Ranade, and E. Koglin, J. Raman Specfrosc., 1977,6, 209.
108 W. Krasser and A.J. Renouprez, in 'Roc. 6th Intern. Conf. Raman Spectrosc.', eds E.D.
Schmid. R.S.Krishnan, W. Kiefer, and H.W. Schrotter. Heyden. London, 1978, p. 522.
109 W. Krasser and A.J. Renouprez, J. Raman Specfrosc., 1979,8, 92.
110 W. Krasser, A. Fadini, E. Rozemuller, and A.J. Renouprez, J. Mol. Sfrucf., 1980,66, 135.
111 W. Krasser, A. Fadini, and A.J. Renouprez, J. Cafal, 1980,62, 94.
112 W. Krasser, H. Ervens, A. Fadini, and A.J. Renouprez, J. Raman Specfrosc., 1980,9,80.
113 R.P. Cooney and P. Tsai, J. Raman Spectrosc., 1980,9,33.
114 H. Yamada and Y . Yamamoto, J. Chem. SOC.,Faraday Trans. I , 1979,75, 1215.
115 R. Rossetti, S.M. Beck, and L.E. Brus, J. Am. Chem. SOC.,1982, 104,7322.
116 R. Rossetti, S.M. Beck, and L.E. BNS, J. Am. Chem. SOC.,1984, 106, 980.
117 R. Rossetti and L.E. Brus, J. Am. Chem. SOC.,1984,106,4336.
118 C.P. Cheng and G.L. Schrader, Specrrosc. Left.. 1979, 12, 857.
119 G.L. Schrader and C.P. Cheng, J. Phys. Chem., 1983, 87, 3675.
120 Y. Soma, M. Soma, and I. Hiuada, J. Phys. Chem., 1985,89,738.
121 K.J. Chao, T.C. Tasi, M.S. Chen, and I. Wang, J. Chem.Soc., Faraday Trans. I, 1981,
77, 547.
View Online
Applications of Raman Spectroscopy to Heterogeneous Catalysis 153

122 C.V. Perrker, W. Pliz, B. Fahlke, E. Loffler, J. Richter-Mendanet, and W. Schirmer, Z.


Phys. Chem., (Leipzig), 1985, 266, 74.
123 A. Campion and W.H. Woodruff, Anal. Chem., 1987,59, 1299A.
Published on 31 October 2007 on http://pubs.rsc.org | doi:10.1039/9781847553225-00102

124 M. Meier, K.T.Canon, W. Fluhr, and A. Wokaun, Appl. Spectrosc., 1988,42, 1066.
125 A. Campion, J . Vac. Sci. Technol. B., 1985, 3, 1404.
126 A. Campion, Ann. Rev. Phys. Chem., 1985, 36, 549.
127 M. Fleischman, P.J. Hendra, and A.J. McQuillan, Chem. Phys. Lett.,1974, 26, 163.
128 D.J. Jeanmarie and R.P. Van Duyne, J . Electroanal. Chem., 1977,84, 1.
129 W.L. Parker, R.M. Hexter, and A.R. S i d e , Chem. Phys. Lett., 1984, 107, 96.
130 A. Wokaun, A. Baiker, W. Fluhr. M. Meier, and S.K.Miller, J. Vac. Sci, Technol. B ,
1985, 3, 1397.
Downloaded by McGill University on 26 December 2012

S-ar putea să vă placă și