You are on page 1of 23

/' MICROPOROUS

MATERIALS
ELSEVIER Microporous Materials 6 (1996) 235-257

Review

X-Ray photoelectron spectroscopy on zeolites and related


materials I
Michael St6cker
SINTEF Chemistry in Oslo, Department of Hydrocarbon Process Chemistry, P.O. Box 124 Blindern, N-0314 Oslo,
Norway
Received 3 January 1996; accepted 5 March 1996

Abstract
XPS has become an important tool for the characterization of surface properties of zeolites and related materials.
Since catalysis is mainly a surface phenomenon detailed knowledge of the behavior of zeolites and related materials
is valuable with respect to the application of those materials in connection with catalyzed chemical reactions and
processes. After a short introduction of the basic principles of XPS some general aspects are addressed before the
more specific items related to XPS investigations of zeolites and related materials are discussed. These chapters are
followed by the characteristics of zeolites and related materials studied by XPS in relation to the surface properties,
the nature and electronic state of metals, their redox behavior as well as investigations of metal clusters/complexes in
zeolites and related materials. Furthermore, the dispersion of metals and metal oxides, the application in catalysis,
the coke formation and finally the acidity and basicity of zeolites will be addressed.

Keywords." X-Ray photoelectron spectroscopy; Zeolites; Molecular sieves; Review

Contents 5. T h e characteristics o f zeolites a n d related


1. I n t r o d u c t i o n ............................................. 236 m a t e r i a l s including i n c o r p o r a t e d species .. 243
2. Physical principle o f X P S ........................ 236 5.1. Surface p r o p e r t i e s ............................. 243
3. X P S - general aspects .............................. 238 5.2. N a t u r e a n d electronic state o f
3.1. X P S intensities a n d s a m p l e m e t a l s / r e d o x b e h a v i o r ....................... 246
c o m p o s i t i o n / q u a n t i f i c a t i o n ............... 238 5.3. M e t a l clusters/complexes in zeolites
3.2. X P S b i n d i n g energies a n d o x i d a t i o n a n d related m a t e r i a l s ........................ 249
states/chemical shifts ......................... 238 5.4. D i s p e r s i o n o f m e t a l s a n d m e t a l
3.3. S h a k e - u p a n d s h a k e - o f f excitations .. 239 oxides in zeolites ............................... 251
3.4. C o m p o s i t i o n - d e p t h profiling ............ 239 5.5. Catalysis a n d c o k e f o r m a t i o n ........... 251
3.5. S t a t e - o f - t h e - a r t X P S ......................... 239 5.6. C h a r a c t e r i z a t i o n o f acidity a n d
3.6. C h a r g i n g / c a l i b r a t i o n ......................... 240 basicity in zeolites ............................. 253
3.7. A u g e r p a r a m e t e r ( A P ) ...................... 240 6. C o n c l u d i n g r e m a r k s ................................. 255
4. X P S a p p l i e d to zeolites a n d related A c k n o w l e d g e m e n t s ....................................... 255
m a t e r i a l s - o v e r r i d i n g i m p o r t a n c e ........... 240 References ..................................................... 255

1 Dedicated to Dr. Hellmut G. Karge on the occasion of his


65th birthday.

0927-6513/96/$32.00 © 1996 Elsevier Science B.V. All rights reserved


PII S0927-6513 (96) 00034-X
236 M. St6cker/Microporous Materials 6 (1996) 235-257

1. Introduction ated by a source of low-energy X-rays under ultra-


high vacuum conditions. Photoionization then
Among the methods of surface analysis based takes place in the sample surface (depth of analysis
on ultra high vacuum techniques, which have been between 1.5 and 6nm). The resultant photo-
developed dramatically during the past years, only electrons (emitted particles) have a kinetic energy
X-ray photoelectron spectroscopy (XPS or ESCA, (Ek) in the range of 20 to 2000 eV, which is related
which stands for 'electron spectroscopy for chemi- [2] to the X-ray energy (hv) and Eb by Eq. 1 (cf.
cal analysis'), besides Auger electron spectroscopy Fig. 1):
(AES) and secondary ion mass spectrometry
Ek=hv- Eb--~p (1)
(SIMS) became well established in the field of
materials surface analysis [1]. XPS provides spa- where h is Planck's constant, v is the frequency of
tially resolved compositional information in the the exciting radiation and (p is the work function
form of spectra from defined areas (microanalysis), of the spectrometer (i.e., the energy required to
and the use of the technique to derive functional bring an electron from the Fermi level of the
group information from chemical shifts in photo- spectrometer to its zero level).
electron energies is well known [2]. There is now If the photoelectrons have sufficient kinetic
a huge knowledge of experience and data in almost energy they are able to escape from the surface by
all areas of material science [2], and the objective overcoming the work function, and photoemission
of this review is to survey XPS results on zeolites happens (Fig. 1). Since the energy levels are quan-
and related materials, including other microporous tized, the photoelectrons have a kinetic energy
materials such as A1PO4 molecular sieves. distribution consisting of a series of discrete bands
which essentially reflect the 'shell' form of
electronic structure of the atoms in the sample [2]
2. Physical principle of XPS (Fig. 2). In a way, XPS is the experimental deter-

From the early pioneering work of Kai Siegbahn Photoelectron(e~,)


KL~ Augerelectron
(Nobel prize awardee 1981) and his group at Levels [ EK (eL)
Uppsala, Sweden, XPS became one of the most
powerful and valuable methods of surface analysis vacuum i~
currently available, mainly through intensive years Fermi
of instrumental, interpretational and applicational
development over the last decade [2]. The physical
2p~(L3~ -=
principles of XPS are well known, and only the 2p~a(!_z) -
essential points will be presented in the following. 2s (L,) ; ~.
Those readers who are interested in more details hw
should consult Refs. [1-4].
XPS has its origin in the investigations of the
photoelectric effect in which X-rays were used as
the exciting photon source. In addition to the Is (K) Photoemis;on Auger process
valence electrons, which provide the bonding for XPS)
the systems, each atom present in the surface Fig. 1. Photoemission (XPS) and the Auger process. Left: An
(except H2 and He) possesses core electrons not incident X-ray photon is absorbed and a photoelectron emitted.
directly involved in the bonding. The so-called Measurement of its kinetic energy allows one to calculate the
'binding energy' (Eb) of each core electron (concep- binding energy of the photoelectron. The atom stays behind as
an unstable ion with a hole in one of the core levels. Right: The
tually, but not strictly, equivalent to the ionization excited ion relaxes by filling the core hole with an electron from
energy of that electron) is characteristic of the a higher shell. The energy released by this transition is taken
individual atom to which it is bound [2]. In the up by another electron, the Auger electron, which leaves the
basic XPS experiment, the sample surface is irradi- samples with an element-specific kinetic energy [4].
M. St~cker/Microporous Materials 6 (1996) 235-257 237

xl0 s Table 1
1.00 | Core-level electron binding energies (in eV) of elements relevant
0"901- 01s to zeolites and related materials (2)
0.801 -
Element lsl/2 2sl/2 2pl/2 2p3/2 3sl/2 3Pl/2 3p3/2
0.70 F
OKL2,aL2,3(Auger)
,~~'°~°I~
"~ 0.601- Nals
~, K4M2,3'~,,31 I
3 Li
4 Be
55
111
0.40 5B 188
8 0.30
0.20 -- "- " ' - ' - ' ~
~ aK/~.3L=,3(Auger)
. - K2~ AI2s
6C
7N
284
399
[ ~,s t v Si2s~,~i2p 80 532 24
0.10 9F 686 31
i i i l i i i - - - llNa 1072 63 31
0.00 1200 1000 800 600 400 200 0
12Mg 1305 89 52
Binding Energy (eV)
13A1 118 74 73
Fig. 2. A1 Kct-excited photoelectron spectrum of 14 Si 149 100 99
offretite/erionite (30% erionite). 15 P 189 136 135
19K 377 297 294 34
20 Ca 438 350 347 44 26
22 Ti 564 461 455 59 34
mination of the kinetic energy distribution by 23 V 628 520 513 66 38
analysis of the photoelectrons produced by expo- 26 Fe 846 723 710 95 56
sure to X-rays. 27 Co 926 794 779 101 60
Routinely used X-ray sources are Mg K~ 31Ga 1298 1t43 1116 158 107 103
(1253.6eV) and A1 K~ (1486.3eV) as incident
particles, which normally cover an area of many
mm 2. A1 K~ radiation can be monochromated, why X-ray sources often contain a dual anode of
and this leads to some focusing, maybe to a spot Mg and AI [2,4].
diameter of about 100/~m (lateral resolution). The Photoelectron peaks are labeled according to
XPS spectrum is usually a plot of the intensity of the quantum number of the level from which the
photoelectrons versus the binding energy Eb. Fig. 2 electron originates. Table 2 summarizes the spec-
shows the XPS spectrum of offretite/erionite. Peaks troscopic nomenclature.
due to Si, A1, K, Na, O and C, due to the
ubiquitous contamination by hydrocarbons, are Table 2
Spectroscopic nomenclature in XPS (4)
easily assigned by consulting binding energy tables
(Table 1). n" lb jc X-ray level Electron level
In addition to the expected photoelectron sig-
nals, the spectrum in Fig. 2 contains peaks due to 1 0 1/2 K ls
Auger electrons as well. The latter arise from the 2 0 1/2 L1 2s
2 1 1/2 L2 2pl/2
deexcitation of the photoion by an Auger trans- 2 1 3/2 L3 2p3/2
ition (Fig. 1), or, in other words, the photoioniza- 3 0 1/2 M1 3s
tion produces not only photoelectrons, but also, 3 1 1/2 M2 3Pl/2
via Auger decay of the electron hole formed, Auger 3 1 3/2 Ma 3p3/2
electrons [2,4]. Obviously, the Auger electron 3 2 3/2 M4 3da/2
3 2 5/2 M5 3d~/2
energy is independent of the X-ray energy used to 4 3 5/2 N6 4f5/2
create the hole, whereas the photoelectron energy 4 3 7/2 N7 4f7/2
is related to the X-ray energy via Eq. 1. Hence,
the apparent binding energy of Auger peaks "Quantum number of the level from which the electron
originates.
appears to change on going from A1 K~ to Mg b Orbital momentum (0, 1, 2, 3.... indicated as s, p, d, f, ...).
Kot radiation, whereas photoelectron peaks do not ¢ Total momentum j = 1 + s (with spin momentum s = + 1/2 or
shift in binding energy. This is the main reason - 1/2).
238 M. St6cker/Microporous Materials 6 (1996) 235-257

3. XPS - general aspects deeper into the sample than the O 1s peak. When
determining concentrations, this effect must be
XPS yields information on the elemental com- taken into account [4]. The quantification of XPS
position (elemental analysis), the oxidation state spectra can be described as follows: The obtained
of the elements and in favorable cases on the value for the peak intensity (peak area followed
dispersion of one phase over another [4]. XPS can by background removal) of the principle spectral
be used to give reasonable quantitative accuracy feature of each element detected in the XPS
(better than ___10%) provided that good calibration spectrum is used to calculate the fractional atomic
standards are used [2]. The detection limit is about concentration, CA of element A by:
10 -3"
A major advantage of the technique is that the IA/SA (2)
cA- (,./s.)
photoelectron energy is dependent on the precise
n
chemical configuration of the surface atoms, and
pronounced chemical shifts are produced in the where I, is the measured peak intensity for element
position of the peaks in the XPS spectrum. XPS n and S, is the relative atomic sensitivity factor
is amenable to virtually all vacuum-compatible (ASF) for that peak.
samples since the incident X-rays do not normally
cause surface damage due to the beam. Therefore, 3.2. XPS binding energies and oxidation
the technique can be used even with very delicate states/chemical shifts
materials [2]. Usually, the sample charging is
minimal. The collection of spectra is normally fast Binding energies are not only element specific
(typically less than 5 min), and the reproducibility but contain chemical information as well, because
of results is high [2]. XPS shows a narrow range the energy levels of core electrons depend slightly
of sensitivities, and the variation in sensitivity on the chemical state/environment of the atom.
across the element range is about a factor of 10. Siegbahn found that this chemical shift is detectable
However, the capability for small area analysis is within a range of 0.1 and 10 eV in magnitude. It
limited, and the depth resolution in composition-- was this ability of the technique that led Siegbahn
depth profiling is poor. Finally, large user data to coin the term 'electron spectroscopy for chemi-
bases and good support data are available, which cal analysis' (ESCA), a term that has often been
make the application of XPS a routine technique. used since, in preference to the more correct name
XPS is compatible with AES and SIMS within the 'X-ray photoelectron spectroscopy' (XPS). In
same instrument, and multi-technique instruments simple terms, chemical shifts arise from the varia-
will continue to have an important part to play in tion of electrostatic screening experienced by core
surface analysis [2]. electrons, as valence electrons are drawn towards
A few (and most important) aspects related to or away from the atom of interest [2]. Usually,
XPS are discussed in the following. the binding energy increases with increasing oxida-
tion state, and for a fixed oxidation state with the
3.1. XPS intensities and sample electronegativity of the ligands. There are excep-
composition~quantification tions, however, as with the alkali metals [4].
As already mentioned earlier, the binding energy
A set of binding energies is characteristic for an is not equal to the energy of the orbital from
element, and XPS can, therefore, be used to ana- which the photoelectron is emitted; the difference
lyze the composition of samples. According to the is caused by the reorganization of the remaining
inelastic mean free path of the elements, the prob- electrons when an electron is removed from an
ing depth of XPS varies between 1.5 and 6 nm, inner shell. As a consequence, the binding energy
depending on the kinetic energy of the photo- of a photoelectron contains both information on
electron. For example, the A1 2p peak in the the state of the atom before photoionization (the
spectrum of offretite/erionite (Fig. 2) probes initial state) and on the core-ionized atom left
M. Strcker/Microporous Materials 6 (1996) 235-257 239

behind after the emission of an electron (the final structure depends on the environment of the atom.
state). Fortunately, it is often correct to interpret One result is that the XPS peaks of metals are
binding energy shifts in terms of the initial state asymmetrically broadened towards higher binding
effects. The charge potential model elegantly energies, and the effect is most pronounced in
explains the physics behind such binding energy d-metals [4].
shifts [4], by means of Eq. 3:
3.4. Composition-depth profiling
Eb i = kqi + q~ +Eb r°f (3)
rij It is (in principle) possible to use XPS to investi-
gate the variation of composition with depth into
in which Ub is the binding energy of an electron the sample in several ways, thereby producing a
from the atom i, qi is the charge on the atom, k is composition-depth profile of the surface of inter-
a constant, qj is the charge on a neighboring atom est. Generally, different approaches are available
j, rij is the distance between atom i and atom j, to arrive at the desired profile, such as analysis by
and Ubef is the suitable energy reference [4]. using different core-levels of the same element, use
The first term in Eq. 3 indicates that the binding of different X-ray energies, use of different emis-
energy goes up with increasing positive charge on sion angles and the so-called sputter-depth profil-
the atom from which the photoelectron originates. ing method (for further details see Ref. [4]).
In ionic solids, the second term counteracts the However, as mentioned previously, the depth reso-
first, because the charge on a neighboring atom lution in composition-depth profiling is rather
will have the opposite sign. Because of its similarity poor.
to the lattice potential in ionic solids, the second
term is often referred to as the Madelung poten- 3.5. State-of-the-art X P S
tial [4].
A few words about energy referencing are appro- Until recently, spatial (lateral) resolution in
priate: Fermi levels of conducting samples which XPS has been of the order of several millimeters,
are in contact are the same. Hence, the Fermi level and this has proven a major limitation and handi-
provides a convenient energy zero. Of course, cap in many areas of application. Within the last
electrons at the Fermi level are still bound to the few years, however, attempts have been made to
metal and have an ionization potential equal to improve spatial resolution both by focussing the
the work function ~0. However, electrons coming incident X-rays and by imaging the emitted photo-
from the Fermi level of the sample are detected electrons. Commercially available small-spot X-ray
with a kinetic energy equal to hv-~o, where ~0 is source photoelectron spectrometers make use of
the work function of the spectrometer [4]. X-ray back-diffraction from a suitable quartz crys-
tal microfocus monochromator. However, the con-
3.3. Shake-up and shake-off excitations ventional (standard) X-ray monochromators are
based on the so-called Johann approximate focus-
XPS spectra of transition metal containing com- sing geometry, which gives excellent monochro-
pounds often show additional peaks in the matization of the source, but produces a fairly
spectrum. These so-called shake-up and shake-off large irradiated area on the specimen. The best
peaks are final state effects which arise when the spatial resolution that can be expected using this
photoelectron imparts energy to another electron method is in the range of 100/~m [2]. In addition,
of the atom. This electron ends up in a higher angle-depending XPS measurements on single crys-
unoccupied state (shake-up) or in an unbound tals contain structural information, such as layer
state (shake-off). As a consequence, the photo- thickness and adsorption geometries [4]. In conclu-
electron loses kinetic energy and appears at a sion, XPS is among the most frequently used
higher binding energy in the spectrum. Those techniques in catalysis. The advantages of XPS are
peaks have diagnostic value, as the precise loss that it readily provides the composition of the
240 M. Stdcker/Microporous Materials 6 (1996) 235-257

surface region and that is distinguishes between and therefore a modified Auger parameter, ~' was
chemical states of one element. Finally, imaging introduced [3]:
XPS is likely to be one of the most fruitful and
~' = ~ + hv = Ek(jkl) + Eb(/) (6)
exciting areas of development over the next few
years [2,4]. The Auger parameter is an empirical measure with
a unique value for each chemical state and can
3.6. Charging~calibration therefore be used as a fingerprint for characterizing
such states [3].
By convention, binding energies refer to core-
level energy relative to the spectrometer Fermi-
level. For conducting specimens in good electrical 4. XPS applied to zeolites and related materials -
contact with the spectrometer, the Fermi levels of overriding importance
both the sample and the spectrometer will be
equivalent, and absolute measurements of core- During the last 20 years, the application of XPS
level binding energies can be made [2]. However, to zeolites and related materials has addressed a
an experimental problem in XPS is that electrically number of questions, like surface sensitivity, the
insulating samples (like zeolites) may charge capability of giving direct information on the
during measurements, since photoelectrons leave valence states, the effective charges, the bond ionic-
the sample. Due to the positive charge on the ities and the band energy structures of zeolites and
sample, all XPS peaks in the spectrum shift by the related materials. Pioneering studies on zeolites by
same amount to higher binding energies. XPS performed by Shpiro et al. [7] and Barr and
Therefore, it is not possible to make absolute coworkers [8-11] were mainly dealing with the
binding energy measurements, and it is necessary variations of the chemical compositions between
to rely on some internal standard for energy refer- the surface and the bulk of the zeolite crystal, the
encing. This calibration is done by using the bind- establishment of the nature of the cation-zeolite
ing energy of a known compound, usually the framework bond, the demonstration of regularities
always present carbon contamination with a C ls as a function of the valence state of transition
binding energy of 284.4 eV [2,4]. elements and their migration to the external sur-
face, the validity of a correlation between the XPS
3.7. Auger parameter (AP) data and those obtained from zeolite catalytic
studies, the states of metals and ligands in com-
The Auger parameter (AP) concept was intro- plexes synthesized in a zeolite matrix, the inter-
duced by Wagner [5,6] to exploit the greater action of highly dispersed metal particles with the
information available from the study of both zeolite framework and studies of adsorption on
Auger and photoelectron lines together rather than zeolites [7]. A number of important reviews by
either alone. The Auger parameter was defined those groups, covering the XPS research on zeolites
originally as performed during the 1970's and 1980's, are well
known, and the interested reader should consult
=Ek(jkl)--Ek(i) (4)
those publications for further details from that
where Ek(jkl) is the kinetic energy of the Auger time [7-11]. At this stage, only a few highlights
transitionjkl and Ek (i) is the kinetic energy of the from that period will illustrate the practical impor-
photoelectron emitted from atomic level i. Eq. 4 is tance of XPS related to zeolites and microporous
equivalent to materials.
The ratios of the surface Si and A1 concen-
e = Ek(jkl) + E b q ) - h v (5)
trations vs. those in the bulk have been found to
where Eb(i) is the binding energy of an electron in be linear. However, the surface and bulk composi-
level i and hv is the energy of the exciting radiation. tions of zeolites of types A, X, Y and mordenite
Eqs. (4) and (5) can produce negative values of c~ are close together, whereas in dealuminated faujas-
M. Strcker/Microporous Materials 6 (1996) 235-257 241

ite the surface Si/A1 ratio increases by a factor of domain of XPS since then [12]. Again, a few
two [7]. This result may lead to a non-uniform recent papers of more general character are
aluminum distribution between the surface and the reviewed to begin with, before the main chapters
bulk crystal. As the binding energy shifts observed dealing with the more specific items related to XPS
in XPS are a direct consequence of variations in and zeolites/micro- and mesoporous materials will
the total charge fields, the following pattern was be discussed.
observed for zeolites: a shift to higher binding XPS binding energies have been measured for a
energy for all constituents, as the size of the cage, number of zeolites (A, X, Y, mordenite and
the cation charge and the Si/A1 ratio increase [8]. ZSM-5) in a wide range of Si/A1 ratios from 1 to
In addition, a collective pattern was realized by 15 in the Na and H forms. Typical trends of the
different binding energies, suggesting the registra- binding energies with the chemical compositions
tion of selective 'groups' rather than 'elemental' were reproduced as known. Measurements of the
chemical shifts. The two primary chemical groups Auger parameter for zeolites of different Si/A1
identified in many of the zeolites seem to be the ratios showed final-state effects to be irrelevant for
SiO2 unit and another one that mimics the the explanation of the binding energy trends.
cation-AIO2 unit [10]. Furthermore, it was con- Furthermore, the Madelung energies at the atomic
cluded that the surfaces of certain zeolites are sites in the zeolite lattice were estimated by lattice-
compromised such that substantial proportions of energy calculations for energy-minimized zeolite
their total surface aluminum are actually present models ranging from Na-X to Na-ZSM-5 and for
in alumina or sodium aluminate residues [10]. models of H-Y of different Si/A1 ratios. These
A first detailed XPS study of the valence band calculations show that the binding energy trends
spectra of zeolites was presented by Barr and upon variation of the A1 content are mainly due
coworkers [11 ]. Based on the general features of to changes in the Madelung potential. Binding
these spectra, the authors subdivided the zeolites energy shifts associated with the exchange of Na
into two, somewhat artificial, classes: group I, by H reflect increased covalent interactions, which
systems of Si/A1 ratios from 1 to 3, typified by reduce the charges on the atoms contributing to
structures dominated by interlaced tetrahedral the Madelung potential. It was suggested that
silica aluminate cages, e.g. A, X, Y and L zeolites, covalent effects may be involved also in binding
and group II, systems of large Si/A1 ratios from 5 energy shifts in the framework under the influence
to >100, typified by structures dominated by of Li and other cations of high electronegativity
interlaced chains, e.g. mordenite, ZSM-5 and sili- [12].
calite [ 11 ]. The nature of the bonding chemistry in zeolites
The valence band spectra exhibited by these has been investigated by Barr [ 13], both by analyz-
zeolites are not just additive renditions of that for ing the XPS core level and valence band results.
silica and for alumina, but rather the zeolite spectra The emphasis in this study was put on the
reveal complex shiftings and alterations. In addi- covalent/ionic character of the chemical bondings.
tion, the valence band results for the group II The zeolite binding energies were seen to shift in
systems appear to be quite similar in appearance a regular progressive fashion with changes in the
to silica, suggesting a silica-dominated system with Si/A1 ratio: higher Eb values with increasing Si/A1
increasing aluminate perturbation as the Si/A1 ratios. The most unusual feature of these changes
ratio decreases, i.e., silicalite < ZSM-5 < is that the binding energies of all the elements shift
mordenite [ 11 ]. in the same direction at the same time. The
Whereas earlier work related to XPS on zeolites following relations were established, based on XPS
was mainly devoted to possible variations of the results, and they are, therefore, particular relevant
chemical composition between the surface and the to the surface regions of the zeolites (however,
bulk of the zeolite crystal, the investigation of they should at least correlate with similar proper-
components introduced into the zeolite lattice or ties for the bulk): The Si-O bond in the silica
deposited on the external surface has been the system is much more covalent than is the A1-O
242 M. St6cker/Microporous Materials 6 (1996) 235-257

bond in an alumina system. Increasing the Si/AI many zeolites are often bound to selected aluminas
ratio provides a system in which the natural cova- or clays to enhance their stability. XPS analysis
lency of the Si-O units in the silica begins to has demonstrated that these binders often coat
dominate, driving up the relative ionicity of the most of the zeolite particles with a thin layer that
AI-O bond and the Na-O bond. The relative is not detected by NMR. In some cases, these
Brensted acidity of these zeolite systems was shown coatings are found to alter various aspects of the
to follow a similar progression. Thus, when the adsorptive and catalytic properties of these sys-
Si/A1 ratio is relatively small, the O - H bond is tems. In other cases, however, N M R and XPS
reported to be relatively covalent and the acidity results suggest detection of the same species [14].
weak. As the relative amount of silica is increased Several zeolites (Na-Y, ZSM-5, L- and T zeolite)
with its covalent Si-O bond, the ionicity (Bronsted and SAPO-5 were investigated by Stoch et al. [15],
acidity) of the O - H bond increases. The correla- resuming that for all network elements only sym-
tion between progressive XPS shifts and the change metrical peaks of the binding energies were
in zeolite bonding properties is also shown to observed, indicating a very homogeneous distribu-
relate to the calculated oxygen charge in a manner tion of the electron densities around the atoms
related to the Sanderson electronegativity (see throughout the lattice. The authors found that the
Fig. 3). The author demonstrated that the oxide binding energies of Si 2p, O ls and A1 2p in
in a zeolite is a highly polarizable unit. Further, zeolites increased with increasing concentration of
he concluded that XPS valence band spectra of aluminum, which is in contradiction to previously
zeolites are often more informative than the core published results. For aluminum, however, the
level results of the same systems [13]. dependence between the A1 2p binding energies
A number of practical aspects related to the and the aluminum content is not as obvious as for
surface behavior of zeolites have been studied by Si 2p and O ls bands. The values either do not
Barr using comparative XPS and N M R results depend on or increase slightly with the aluminum
[14]. During the conversion of parent Y zeolites content. The changes might be interpreted either
to their ultrastable form, certain cation to be due to variations in the polarity of AI-O
by-products have been found by XPS to be selec- bonds and/or to the increasing concentrations of
tively transported to the outer surface. counter cations that localize charges and, hence,
Furthermore, in connection with applications cause a slight depletion of electrons at the A1a ÷
lattice site. The XPS results indicate that phos-
SiO2 phorus (in SAPO-5) increases the electron densities
at all other network elements, leading to a lowering
A of the binding energy especially of oxygen, whereas
532.0 the effect upon aluminum and silicon is less pro-
nounced [ 15 ].
0
C ~ , ~axNaY The electronic structure of zeolites was demon-
Q
CaXo ~
_c strated to depend strongly on both the composition
_~¢ 531.0 and the cation involved. It was suggested by
an
\\.,x Okamoto et al. that the basic strength of frame-
oc. \\ work oxygen increases with increasing A1 content,
regardless of the crystal structure, and with
530.¢
decreasing electronegativity of the counter cation
-0.2 -0.3 -0.4 o Na~120* [16].
No For zeolites that exhibit ion conductivity, the
Fig. 3. Shift in zeolite O ls binding energies with change in
surface charge in the application of surface spectro-
oxygen charge, No, based on Sanderson electronegativity scopy, like XPS, may be significantly reduced by
(reproduced by permission of Elsevier Science B.V., using elevated measurement temperatures. This
Amsterdam). technique permits the straightforward detection of
M. St6cker/Microporous Materials 6 (1996) 235-257 243

differential-charging effects in XPS with zeolites. 0.6

This method is applicable to those zeolites that 0.5 Y(U.S.)


exhibit an increased conductivity at elevated tem-
peratures, however, a limitation lies in the elevated + 0.4
/ Na-A
temperatures required for the measurements,
which may give rise to modifications of the 0.3
samples [ 17]. FerrieriteJ u ZK'5
0.2 -I.-Edonite~ __Clinoptilolite
,,,°,,,e

5. The characteristics of zeolites and related 0.1 ,~o Merdenite


~ZSM-5
materials including incorporated species 0 I I I I I
0.1 0.2 0.3 0.4 0.5 0.6
5.1. Surface properties Framework AI/(Si + AI)

Fig. 5. Plot of framework A1/(Si +A1) vs. surface AI/(Si +A1)


The surface composition and electronic structure (reproduced by permission of Elsevier Science B.V.,
of a number of synthetic and mineral zeolites with Amsterdam).
well characterized bulk and framework composi-
tions over a wide range of Si/A1 ratios have been decreases with an increase in the Si/AI ratio. The
characterized by XPS for the determination of binding-energy separation [Si 2 p - A 1 2s] also
Si/A1 composition of external layers [18]. The shows a similar trend.
results of the XPS and bulk analyses of the zeolites The dealumination of zeolites by different meth-
studied showed that, with the exception of ther- ods and the subsequent migration of the removed
mally treated samples exhibiting significant frame- aluminum species has been monitored by XPS,
work dealumination, the surface Si/A1 ratios are since catalysis is mainly regarded as a surface
generally equal to or slightly greater than the bulk phenomenon, and dealuminated zeolites are still
stoichiometric ratios for most of these zeolites (see of great interest from an industrial point of view
Fig. 4). In addition, the surface composition of due to their modified acidity. Therefore, a number
the same series of zeolites is generally silica- of authors have studied this important topic mainly
enriched relative to the framework composition on Y zeolite but also on other microporous materi-
(except for USH-Y) as determined by 29Si MAS als. However, the results do not always point in
N M R spectroscopy (see Fig. 5) [18]. Furthermore, the same direction. Depending on the dealumina-
the binding energy of the A1 2p photoelectron line tion method used and which binding energies for
Si (Si 2s or Si 2p) are applied for calculating the
0.6 Si/A1 ratio, different conclusions were drawn by
the different groups.
0.5 H'Y(U.S.)~ / The enrichment of aluminum at the surface of
<
0.4 Y zeolite has been observed by several authors
÷
and is interpreted as an accumulation of non-
v
< 0.3 framework aluminum species which migrate out
8 of the micropore system towards the zeolite surface
0.2 Clin°pt~--sFerderite [19-22]. Such a Si/A1 profile indicates that the
e dealumination mechanism is diffusion controlled
0.1
"Chabazite within the zeolite pores [21 ]. The aluminum species
~"ZSM-5
I I I I I expelled from the zeolite lattice and enriching the
0.1 0.2 0.3 0.4 0.5 0.6 zeolite surface most likely form a thin amorphous
Bulk AI/(Sl + AI)
aluminum oxide/hydroxide layer. Secondary ion
Fig. 4. Plot of bulk A1/(Si+A1) vs. surface A1/(Si+A1) (repro- mass spectrometry (SIMS) detects the dealumi-
duced by permission of Elsevier Science B.V., Amsterdam). nated zeolite already at depths of about 50 .A [20].
244 M. Stdcker/Microporous Materials 6 (1996) 235-257

The Si/AI ratios of the external layers show for num to the surface takes place during the deep-
the thermochemically dealuminated Y zeolite a bed calcination of NH4-Y zeolite, the extent of
strong dependence on the temperature of treat- dealumination being higher inside the crystal than
ment, and at high temperatures, can become nearly at the outer surface [30].
three times higher than that of the bulk [19]. A The transport resistance experienced by ethane
quantitative analysis of XPS intensities for dealum- molecules on traversing the surface layer of
inated Y zeolites was done by Grohmann and CaNa-A type zeolite crystallites was examined by
Gross who concluded that the respective A1/Si K~irger et al. [31]. It was found that a hydrother-
XPS intensity ratios of highly dealuminated Y mal pretreatment leads to an enhancement of the
zeolites are explained by dispersion of the non- surface resistance which is most pronounced for
framework A1 on the mesopores surface [23]. the specimen with the lower calcium content.
However, partial removal of aluminum from the Studying the surface composition of the zeolite
surface of zeolites (Y, X and A) during a dealumi- crystallites by XPS, the enhancement of the surface
nation procedure was observed by several groups barrier was shown to be accompanied by a decrease
[24-26]. In addition, Si enrichment was investi- of the cation content in the surface layer, indicating
gated in detail by Japanese researchers using XPS a structural collapse of the surface layer of the
with argon-ion etching, who concluded that sili- zeolite crystallite [31 ].
con-rich configurations are localized at the surface The composition of the surface layer of different
rather than in the interior or bulk of Y zeolite or Fe-, Co- and Cu-doped A zeolites have been
mordenite (see Fig. 6) [27]. Finally, the external investigated by Knecht et al. applying XPS. They
shell formed on a Y zeolite during steaming was
concluded that the concentrations of the transition
shown by XPS to contain both aluminum and
metals in the surface layer and the bulk material
silicon extracted from the framework [28]. The
are significantly different, with higher transition
presence of tricoordinated aluminum at the surface
metal concentrations in the surface [32-34].
of dealuminated mordenites with bulk Si/A1 atomic
Framework silicates with the sodalite structure
ratios higher than 40 were registered by Remy
have been studied by Herreros et al. [35], who
et al. [29].
demonstrated that purely siliceous sodalite gave a
The surface atomic ratio of framework and non-
unique set of binding energies and valence band
framework aluminum in a deep-bed steamed H-Y
patterns similar to those for silica, while the XPS
zeolite and in an H-Y zeolite dealuminated by
SIC14 has been determined by XPS. The results pattern of an aluminosilicate sodalite is similar to
showed that the deep-bed dealuminated zeolite that of zeolite Na-A. The small differences in
contained two aluminum species (75.0 and 73.8 eV, binding energies for aluminosilicate sodalite are
A1 2p transitions), which can be assigned to frame- due to the presence of some K ÷ and Mg 2- and
work and non-framework atoms, respectively. the fact that the framework Si/A1 ratio is slightly
Moreover, migration of the non-framework alumi- higher than unity. XPS was thus shown to be
sensitive to the structure as well as to the composi-
tion [35]. In continuation of this work, solid-state
Surface Interior N M R showed that blue and pink ultramarines
etching) with bulk compositions Si/AI=I.04 and 1.11,
respectively, violate Lrwenstein's rule. XPS
revealed that the surface Si/A1 ratio of blue ultra-
: : ) Bulk
marine is 1.20, which corresponds to positive shifts
in core level binding energies. Pink ultramarine
:i:i:i:i:i:i:i:i:i:i:i:i:i:i:i:i?i:i:!:!:!:!:i:i:i:!:i:i:i:i:i:i:i has an even higher surface Si/A1 ratio and a
Fig. 6. Conceptional illustration of surface, interior and bulk corresponding growth in its core lines. Core level
of a zeolite (reproduced by permission of the American peak positions for zeolite losod, a compact and
Chemical Society, Washington, DC). highly aluminous framework silicate (Si/A1 about
M. Strcker/Microporous Materials 6 (1996) 235-257 245

1.0), are very close to those for aluminosilicate varies from 1711.3 to 1708.8eV. The Si KL2L 3
sodalite and zeolite Na-A [36]. Auger line ranges from 1608.0 to 1610.6 eV. The
XPS measurements were used to identify surface Si 2p binding energy ranges from 101.3 to
properties of silica-rich zeolites such as offretites, 103.5 eV [37].
ZSM-5, ZSM-11 and mordenites. The peak posi- The chemical state of A1 at the surface of
tions of the AI and Si Auger lines were measured dealuminated mordenites has been investigated by
in Mg K~ XPS experiments. The A1 and Si XPS. The A1 2p signals can reasonably be decom-
KL 2L 3 Auger transitions are referenced to the C posed into two components. The A1 KLL Auger
ls line at 284.6eV. Data from the A1 and Si transition is more sensitive to the A1 coordination
KL2L3 Auger lines and the A1 2p and Si 2p than the AI 2p transition, and three components
transitions for the silica-rich zeolites have been can be separated in the corresponding peaks. It
used to construct A1 and Si chemical state plots appeared that three types of A1 were present at
shown in Fig. 7 and Fig. 8, respectively [37]. the surface of dealuminated mordenite, namely
The A1 chemical state plot has a wide range of hexa-, tetra- and tricoordinated A1 [38].
modified Auger parameters from 1456.2 to Two series of offretite and omega zeolites were
1460.3 eV and binding energies ranging from 73.4 modified by ammonium exchange and sequential
to 75.2 eV. The A1 KL 2L 3 transition has a very steaming/acid leaching by Ponthieu and Grange
large range from 1383.2 to 1387.1 eV. The same [39]. XPS was used to quantitatively and qualita-
materials have a much smaller spread on a Si tively analyze the surface of these solids. The
chemical state plot. The modified Auger parameter superficial A1 concentration significantly differs

/ // / o./~/
1387/
/o/ 1460
/ / / / -/
/ / [] 1459
[] / / 0 // /
1386

-: / -

,~'/ /
1388 / []
®

o ,/ / /
.u_
i ~.

x
J
[] / /- 1458
,- Z / /I-,
/ :
I
1384 ~ [] 16o8 ~ 17o9 ~
1487

// /,
/°/ /
74 73
/ / /
76 78 104 103 102 101
2P Bindingenergy (eV) 2P Bindingenergy (eV)

Fig. 7. Aluminum chemical state plot, 1-4: different offretites, Fig. 8. Silicon chemical state plot, 1-4: different offretites, 5:
5: H-ZSM-5, 6: H-ZSM-11, 7: H-MOR, 8: Norton-MOR, 9 H-ZSM-5, 6: H-ZSM-11, 7: H-MOR, 8: Norton-MOR, 9 and
and A: different Na-A, X: Na-X, Y: Na-Y, Z: H-zeolon (repro- A: different Na-A, X: Na-X, Y: Na-Y, Z: H-zeolon (reproduced
duced by permission of the American Chemical Society, by permission of the American Chemical Society,
Washington, DC). Washington, DC ).
246 M. St6cker/Microporous Materials 6 (1996) 235-257

from the bulk one in synthetic solids. The greatest 5.2. Nature and electronic state of metals/redox
modifications arise from acid-leaching treatments. behavior
The surface of hydrothermally dealuminated
offretites often contains relatively more A1 than Transition metal containing zeolites are impor-
does the bulk, whereas dealuminated omega zeo- tant catalysts for a number of chemical reactions.
lites generally display Al-depleted surfaces [39]. The nature and electronic state of those metals
Furthermore, offretites, erionites and their are, therefore, important items to investigate, and
intergrowths (T zeolites) were studied by XPS by a number of papers dealing with this topic are
a few more groups. Surface dealumination was reviewed in the following.
observed for all investigated zeolites [40,41]. XPS studies of hydrated, dehydrated and
ZSM-5 and ZSM-11 samples have been investi- reduced CuNa-Y and CuNa-X zeolites were car-
gated by several groups with respect to the chemi- ried out by Kevan et al. [50,51]. With sufficient
cal compositions at the surface and bulk. XPS analytical resolution, both the Cu 2pl/2 and 2p3/2
data showed that the outermost layers are on an transitions were resolved into two components in
average, richer in A1 than the inner layers [42], hydrated CuNa-Y. The higher binding energy
which was confirmed for ZSM-5 samples treated component (about 935-936 eV) and the shake-up
by strong dealumination procedures, too [43]. satellites decreased in intensity upon dehydration.
The higher binding energy component was
However, an inhomogeneous distribution of alumi-
assigned to octahedrally coordinated Cu 2+ in the
num in ZSM-5 polycrystalline aggregates, con-
large a-cages of the zeolite, and the lower binding
sisting of a siliceous outer surface and an
energy component (about 933-934 eV) to tetrahe-
aluminum-rich interior was found by Hughes et al.
drally coordinated Cu 2+ in the small r-cages of
[44]. The surface depletion of aluminum marks a
the zeolite [50]. Cu + and Cu °, obtained by
sharp contrast with some recent surface studies of
hydrogen reduction at high temperature, showed
ZSM-5. For example, Suib et al. and Dwyer et al.
single XPS peaks at about 932eV without
found no differences between the Si/A1 ratios of
shake-up satellites [51].
the surface and bulk [45,46]. However, von Four states of Cu have been characterized in A,
Ballmoos and Meier [47] measured an enrichment X and Y zeolites with XPS by Sexton et al. [52].
of aluminum at the surface of ZSM-5, while
By applying the 2p3/2 and L 3 M4,5 M4,5 Auger
Derouane et al. [48] found similar surface and lines the authors identified Cu 2+, Cu +, 1 nm metal
bulk Si/AI ratios for large crystals (5-8 #m) but clusters and bulk metal crystallites. The Cu-species
an enrichment of surface aluminum for smaller present in the zeolite depend on the degree of
crystals (< 1 #m). As Derouane et al. suggested, reduction. Cu + can be exclusively generated by
this variance of results is due to the fact that mild reduction of X or Y zeolite with CO. One nm
ZSM-5 is not a uniform material [48]. Hence clusters are generated in the supercages of zeolite
conclusions drawn about the aluminum distribu- A by mild hydrogen or CO reduction [52].
tion of ZSM-5 prepared by one particular method Migration of Cu species into the zeolite channels
may not necessarily be applicable to all ZSM-5 at elevated temperatures was observed for a mix-
preparations [44]. ture of Cu20 and NH4-Y zeolite by Jirka et al.
The XPS spectra of the natural zeolite natrolite [53]. The same group demonstrated that the Cu
were measured, and a comparison of the binding 2pa/2 line-shape changes for cupric ions exchanged
energies of the Si 2p, A1 2p, Fe 2p3/2, Na 2s, Ca in Y-zeolite are caused mainly by their reduction
2p3/2 and O ls transitions with published data on to Cu ÷ during the XPS experiment. Similar (but
similar substances confirmed that the coordination much faster) effects, understandable in the same
numbers of Si and A1 in the natrolite structure way, have been observed for Cu-ZSM-5 zeolite
were equal to 4. In addition, the following coordi- [541.
nation numbers were found: Fe=4, Na=6, C a = The determination of the oxidation state of Cu
6 and O = 2 [49]. in Cu-exchanged ZSM-5 is important for under-
M. St6cker/Microporous Materials 6 (1996)235-257 247

standing its role in the mechanism of decomposi- opposite observation has been reported [59].
tion and selective reduction of NO [55]. Cu-ZSM-5 Nickel in NH4NiNa-Y zeolite has been observed
catalysts prepared by ion exchange and by impreg- in a highly dispersed state [60]. An intrinsic
nation have been investigated by XPS. In both method for measuring the true kinetic and binding
preparations, the external surface region of the energies of photoelectrons has been described by
catalyst is highly enriched in Cu. In the ion- Steinbach et al. [61] applying their technique to
exchanged zeolite, the observation of high XPS Ni containing A, X, Y and ZSM-5 zeolites. The
binding energies and unexpectedly low Auger binding energies of O ls and Si 2p transitions
kinetic energies showed that Cu is present as increased with increasing Si/A1 ratio of the zeolites,
isolated ions or small clusters. In the impregnated for the A1 2s transition, however, the increase was
material large aggregates of metallic Cu at the less marked. The influence of the cation on the
external zeolite surface were produced by reduction binding energy values of the O 1s, Si 2p and AI 2s
in hydrogen, which were not dispersed by treat- was examined only for Y zeolite: no influence of
ment in oxygen or NO. In the exchanged materials, the cation on Eb was observed [61].
the Cu became more evenly distributed across the The behavior of Fe 3÷ cations in H-Y zeolite
zeolite crystal as a result of calcination or reductive which had been subjected to various treatments
treatment [56]. has been studied by XPS. The self-reduction of
The surface characteristics of Ni containing Fe 3+ to Fe 2÷ after vacuum heat treatment and
ZSM-5 prepared by ion exchange, impregnation the partial reduction to Fe ° after hydrogen treat-
and mechanical mixing have been monitored by ment was found in the zeolite surface layers in
XPS. Ni can exist in many forms on the surface accordance with results of studies of the bulk
of the zeolite, e.g. as NiO, Ni(OH)2 and zeolite. The binding energies of Fe cations were
NiA1204. The degree of reduction of Ni 2÷ species found to be close to those of the corresponding
depends mainly on its form, and decreases in the oxides [62]. Furthermore, the external surface
order NiO>Ni(OH)2>NiA1204. The surface composition of Fe-exchanged Y zeolite with
analysis indicates that ZSM-5 does not have a respect to percent exchange and thermal treatment
uniform composition. Surface enrichment of alu- was investigated by Kulkarni et al. [63] who
minum and nickel was observed and can be concluded that the Si/A1 ratios of the external
explained be several factors, including the sample surface were higher than the corresponding bulk
preparation method, the stage of the redox cycle ratios and showed a small increase with the percen-
and the treatment conditions [57]. A Ni-exchanged tage ferric exchange. This indicates that the exter-
ZSM-5 sample was used for the xylene isomeriza- nal zeolite surface is aluminum deficient [63].
tion and checked with respect to the state of the The extent of the electron deficiency of Pd
nickel under those operating conditions. The clusters in Y zeolite with varying proton concen-
authors found that the presence of nickel sup- trations and locations has been identified by XPS.
pressed catalyst aging significantly and, in addi- The electron sharing between Pd clusters and
tion, the nickel ions did not undergo any major protons in Pd/Na-Y zeolite results in an upward
chemical modification during the pilot plant trial shift of the binding energy of Pd 3d5/2 of about
[58]. 0.4 eV. No surface enrichment of Pd was observed,
XPS analyses of hydrated and hydrogen-reduced and the metal was homogeneously distributed.
NiNa-X and NiCa-X zeolites have been per- Reduction of the ion-exchanged Pd/Na-Y zeolite
formed, indicating that reduction at 327°C is in hydrogen at elevated temperatures led to a Pd
sufficient to reduce some of the Ni 2÷ in NiNa-X 3d line shift to lower binding energies (see Fig. 9)
to the metallic form. Deconvolution of the broad [64].
XPS peak shape indicates three components XPS studies of Pd-exchanged NaX zeolite indi-
assignable to Ni 2÷, Ni ÷ and Ni °. XPS data indi- cate that, under suitable ion exchange and sample
cate that hydrogen reduction of Ni 2÷ to Ni ÷ and treatment condition, unstable oxidation states of
Ni ° occurs at lower temperatures in NiNa-X than Pd 3+ and Pd ÷ can be stabilized and characterized.
in NiCa-X zeolite. Concerning the Y zeolite, the The peak at a binding energy of 339.0 eV (Pd
248 M. Sti~cker/Microporous Materials 6 (1996) 235-257

the ion-exchanged and impregnated samples


revealed that transition metal ions are really
exchanged into the zeolite but their distribution is
not homogeneous in the subsurface layer. The
simultaneous presence of Co 2÷ and Pt 2+ ions
450
results in an increased sensitivity of Pt to oxidation
or in the formation of a mixed Pt oxide/ CoOn
350' phase. Co 2÷ ions also affect the electronic state
of Pt o particles as displayed in the positive binding
300 energy shift of the Pt 4d5/2 transition. The binding
energy shift of the Co 2p3/2 line evidences the
250 formation of bimetallic particles within the zeolite
support. Pt 2÷ ions exchanged before the Co 2÷
200' ions hinder the migration of the latter into deeper
zeolite layers, suggesting the formation of 'cherry'-
150' type particles. Pt ° particles were unambiguously
proven to assist the reduction of Co 2÷ ions in the
100 ' cavities of Na-Y zeolite [67].
Rhodium exchanged Y zeolites have been shown
20
to exhibit a wide variety of catalytic properties,
337.0 347.0 and XPS studies indicate that Rh a ÷ may be ther-
Ee'eV
mally reduced to Rh metal in the zeolite Y lattice.
Fig. 9. Positions of the Pd 3d line at increasing temperatures It was strongly suggested that metallic Rh pro-
for Pd/Na-Y zeolite (reproduced by permission of the Royal duced by the activation of Rha+-exchanged Y
Society of Chemistry, Cambridge.). zeolite rather than the Rh 3÷ cations, are the
probable catalytic species in so-called Rh 3÷ zeolite
3d5/2 transition) in oxidized samples of Pd/Na-X catalysts [68]. A stable zeolite-supported Rh cata-
zeolite can be unambiguously assigned to Pd 3+. lyst was studied by Liu et al., and the active Rh
In samples reduced in static hydrogen the XPS component was found to be trivalent, and Rh in
peak at 336.4 eV corresponds to Pd + , but if flow- the catalyst was in form of ions which partly
ing hydrogen is used this peak seems to correspond occupied the Na ÷ sites in the lattice of ZSM-5 [69].
to small charged Pd clusters [65]. XPS has been Ru-exchanged Na-Y zeolites were studied with
used to characterize different oxidation states of XPS in order to determine changes in the location
Pt in Y zeolite. Monoatomically dispersed Pt ° and state of the Ru after subjecting the samples
atoms exhibit an usually small relaxation energy to both oxidizing and reducing environments.
(+ 1.3 eV shift), and 10 or 20 ,~ metallic Pt clusters When oxygen was excluded from the system, Ru
present a significant positive XPS shift (+ 0.7 eV). remained inside the zeolite cavities following the
A comparison with Pt clusters on silica (+ 0.3 eV) reduction in hydrogen and after the zeolite had
and a study of the ability of charge transfer catalyzed the methanation reaction. At elevated
complex formation have shown that electron temperatures, before or after reduction, the pres-
transfer occurs between the metallic clusters and ence of oxygen resulted in the formation of
the zeolitic support, resulting in strong electron RuO2 on the external surface of the zeolite. The
donor properties of the support [66]. Under the peak intensity of the Ru 3d5/2 line increased by
conditions applied for the ion exchange of Na + more than a factor of five when Ru migrated to
by transition-metal cations (Pt 2+ and Co2+), alu- the surface. Positive binding energy shifts of
minum containing ensembles of Na-Y zeolite approximately 1 eV were detected for small clusters
migrate toward the external surface resulting in (< 10 .A) of reduced Ru when compared with larger
surface enrichment of aluminum. Comparison of metal particles within the zeolite or on its external
M. St6cker/Microporous Materials 6 (1996) 235-257 249

surface [70]. A Ru/K-L zeolite has been studied anatase (octahedral). Non-framework TiO2
by Liu et al. [71], who demonstrated that Ru 3+ species containing octahedrally coordinated Ti
and Ru ° atoms are present in the zeolite catalyst. could also be distinguished by XPS [75].
The activity for carbonyl hydrogenation remained Ti-silicalite samples with MEL structure (TS-2)
the same after several reaction runs, independent and different Ti loadings have been examined by
of the conversion [71]. Trong On et al. [76] using XPS (among others).
Faujasite-type zeolites containing La introduced It was shown that as the Ti content increased, the
by various techniques and in different amounts Ti 2p3/2 binding energy decreased from 459.8 eV
were studied by XPS in order to elucidate the to 458.3 eV, which corresponds to a change from
influence of the zeolite matrix on the binding tetrahedrally coordinated Ti to octahedrally coor-
energy and the shape of the La 3d line. Significant dinated Ti-species [76].
differences in the La 3d line shapes and binding A series of alkali-cation-exchanged and some
energies between La203 and La zeolites were proton-exchanged zeolites have been investigated
ascribed to final-state effects arising from the with respect to cation-framework interactions. The
hybridization of La final states with valence bands XPS study revealed that the higher the electronega-
of different extension. The La 3d line shape of the tivity of the counterion is, the lower the framework
La zeolites investigated at elevated temperatures element binding energies of Si 2p, A1 2p and O ls
was found to respond to structural changes in the are. However, the extent of these changes depends
strongly on the Si/A1 ratio of the zeolites [77]. A
surface region of the samples (dehydration, ion
study of Na-, Li- and Ce-exchanged Y-type zeolites
migration, formation of highly dispersed La-O
using conventional Mg K~ and high-energy mono-
structures on extra-framework sites) [72]. Rare-
chromatic Ag L~ radiation has been described by
earth ion-exchanged Y-type zeolites were investi-
Edgell et al. [78]. The chemical changes induced,
gated by Arakawa et al., who showed that the 4f
predominantly in Na-Y zeolites, involve the reduc-
levels of Eu-Y zeolite were drastically changed by
tion of the Na + cations to the metallic state,
degassing at low temperature: the relative inten-
apparently following their diffusion from the bulk
sities of the 4f levels for Eu 3+ decreased, and the to the surface. The Auger parameters of silicon
4f line of Eu 2+ increased as the sample was and aluminum and the derived oxygen polarizabil-
degassed at 400°C [73]. ity decrease with the electronegativity of the
Catalysts consisting of Ga203 mechanically charge-balancing cation for this series of simple
mixed with H-ZSM-5 zeolite have been examined cations [78].
via XPS both before and after reduction with The interaction between the mono- and poly-
hydrogen. The XPS measurements have shown cationic forms of Y-zeolites containing Ni, Co,
that there is a Ga transfer upon hydrogen reduc- Cu, Cr and CO, oxygen or CO +oxygen under
tion of Ga203/H-ZSM-5 catalysts prepared by conditions similar to those of catalytic CO oxida-
mechanical mixing the pure components. tion has been monitored by XPS. Upon interaction
Experiments with Ar+-etched samples show that of Ni-containing Y zeolites with CO or with
Ga ions are transferred into the bulk of the zeolite. CO+oxygen, Ni 2+ is reduced to Ni + and Ni °.
The Ar ÷ bombardment leads to an increase in the The ability of metal reduction decreases in the
Ga 3d binding energy [74]. following series of polycationic forms:
The chemical bonding of Ti in Ti-silicalite has Cr > Co > Ni > Cu. The treatment of CuNiNa-Y
been investigated using XPS. A distinctively higher and CoNiNa-Y zeolites with CO leads to the
binding energy of the Ti 2p photoelectrons in formation of Cu +, Cu ° and Co °, respectively [79].
Ti-silicalite was found compared to those mea-
sured for titania- or anatase-containing silicalite. 5.3. Metal clusters~complexes in zeolites and
Ti had the same oxidation state in all solids related materials
investigated. The difference in binding energy was
attributed to different oxygen coordinations of Ti The position, structure and distribution of metal
in the framework of Ti-silicalite (tetrahedral) and clusters/complexes in zeolites and related materials
250 M. St6cker/Microporous Materials 6 (1996) 235-257

is of importance with respect to catalytic applica- for Si 2s, AI 2p, O ls and Na ls remained
tions of modified zeolites. Therefore, several unchanged, a small shift toward lower energies for
attempts have been made to characterize these the Co 2p and N 1s transitions was observed when
metal-complex catalysts in the light of their poten- CoPc was in contact with the zeolite (see Fig. 10).
tial use. The surface atomic ratio for Co/Si was higher than
An XPS study of Mo clusters in zeolite Y formed for the bulk Co/Si ratio. On the other hand, the
by adsorption and decomposition of Mo(CO)6 has corresponding N/C atomic ratios are about the
been reported by Yong et al. [80]. The XPS data same [84].
showed that this method produced a uniform XPS investigations of methylene blue incorpo-
dispersion of Mo. The Mo clusters have a Mo 3d rated into faujasites and AIPO4 molecular sieves
binding energy ca. 1.2 eV higher than that of bulk were reported by Hoppe et al. [85]. The exchange
Mo, which is attributed to their small size (equal of the methylene blue cation into the zeolites
or less than 12 A) [80]. The adsorption and decom- resulted in high dispersions of the dye, but could
position of the volatile carbonyl complexes cause a local destruction of the zeolite framework
Mo(CO)6, Fe(CO)5, Co2(CO)8 and Co(CO)3NO at the crystal surface. The structure-directing role
in various zeolites have been studied by XPS. of the dye was recognized for methylene blue
Mo(CO)6 produces highly dispersed zero-valent loaded SAPO-5 and SAPO-34, exhibiting an
clusters in the zeolites Na-Y, Na-X and K-L, and improved phase purity [85].
a mixture of oxidized species in La-Y zeolite.
Fe(CO)5 produces initially zero-valent clusters in
Na-Y, but these migrate to the external surface on
heating to higher temperatures. Migration of iron
out of the zeolite is inhibited by exchanging the
divalent cations Co 2÷ and Ni 2÷ into the zeolite.
Coz(CO)a decomposes extensively on the external
surface of Na-Y and Na-X zeolite, but the mono-
nuclear complex Co(CO)3NO is readily adsorbed
into Na-Y; decomposition produces a mixture of
zero-valent and oxidized cobalt species [81 ]. PY
Transition metal complexes fixed in Y zeolites
by 'surface assembling' (Ni-, Co-, Cu- and
Ru-phthalocyanines) have been investigated by
XPS. Metal phthalocyanines were found to be
located in supercages as isolated complexes which laY
n$
interact with zeolitic OH groups via meso N atoms
of the porphyrin ring [82]. Furthermore, XPS was ~aY
used to investigate the formation and location of
Ni 2÷- and C02÷-bis(dimethylglyoximato) com-
plexes following the reaction of Ni 2÷- and
Co 2÷-exchanged Na-X zeolites with dimethylgly-
oxime. The formation of the complexes and the ~e
n$
location within or on the outer surface of the 810 800 790 780 770
zeolite matrix was sensitive to the reaction condi- eV
tions and the sequence of reaction steps [83].
Cobalt phthalocyanine (CoPc) was synthesized Fig. 10. XPS binding energies of the Co 2p transitions for cobalt
phthalocyanine (CoPc), CoPc with Na-Y, CoPcNa-Y, CoNa-Y
within the cavities of Y zeolites and characterized and CoNa-Y with pyridine (py). The vertical scale is not the
in all steps of the synthesis and purification by same for all the samples (reproduced by permission of the
(among others) XPS. While the binding energies American Chemical Society, Washington, DC).
M. Strcker/Microporous Materials 6 (1996) 235-257 251

5.4. Dispersion of metals and metal oxides in isolated Cr 5+ and Cr 6+ ions present in the oxidized
zeolites zeolite bulk were not detected by XPS in the
surface layers because of their reduction in vacuo
Metal loaded zeolites are widely used in petro- and/or by X-ray irradiation in the XPS system
chemical processes and are broadly investigated [89].
with respect to applications within catalysis. The A series of MOO3, NiO and MoO3 +NiO on
zeolite matrix has proved to be especially suited ultrastable H-Y zeolite (USHY) catalysts with
for the production of monomodal metal disper- different oxide loadings and different procedures
sions, which have been identified mainly by TEM. of preparation have been studied by (among
The conclusion has been drawn from the electron others) XPS. It has been found that Mo and Ni
micrographs that monodispersed metal phases can species are highly dispersed, and both oxides can
be located exclusively inside the zeolite matrix. It be found in the supercages of the Y zeolite. In the
is, therefore, desirable to check such an assumption case of N i O + M o O a / U S H Y catalysts, an inter-
by a surface sensitive method like XPS. This was action between Ni and Mo was established through
done by Schulz-Ekloff et al. with a series of oxygen bonds. The degree of interaction depends
faujasite supported catalysts with monomodal dis- on the order of incorporation of the metals and
persed phases of nickel, palladium and platinum, on the Ni/Mo ratio. The strongest interaction was
exhibiting narrow particle size distributions as observed in the sample in which the Mo was
characterized by XPS [86]. impregnated first [90]. The identification and study
No XPS evidence was found for the migration of the electronic properties and location of CdS
of Pt to the external surface of Pt/Ca-Y zeolite and CdO particles in faujasites have been per-
crystallites, whereas this phenomenon was indeed formed using XPS. Depending on the preparation
observed for a sample of Pt/Na-Y which was procedure, CdS aggregates have different disper-
reduced directly in hydrogen at 350°C. In addition, sions and can be located either inside the zeolite
differences in metal-support interactions were not structure or on the external surface. Small CdS
evident from the binding energy data for the particles (1.5-2 nm) exhibit XPS band shifts indi-
several types of reduced platinum examined in this cating a semiconductor size quantization effect.
study [87]. Pt and Pd ions in Y-type zeolites were Larger CdS particles (6-8 nm) exhibit electronic
investigated by Vedrine et al. [88], and both ions and relaxation properties rather close to those of
were found to be ionically bonded to lattice bulk CdS. Strong surface enrichment with these
oxygen. Atomically dispersed Pt ° and Pd ° were particles was also observed. The CdO aggregates
shown to give significant XPS positive chemical are formed during Cd-Y treatment with NaOH
shifts: + 1.3 eV and + 1.4 eV for Pt 4f and Pd 3d and calcination in air. The identification of CdO
lines, respectively. These shifts were assigned to is based on an abnormally low binding energy of
smaller relaxation energies. The positive XPS Cd, which is characteristic for CdO only [91].
chemical shifts detected for metal Pt aggregates on Finally, a paper dealing with the quantification
zeolites compared with those for Pt on silica of XPS data of supported catalysts using the
strongly suggest that electron donation occurs so-called sheet stacking model with surface segre-
from the metal aggregates to the zeolite lattice [88]. gation was published by Sayari et al., and the
XPS has been used to elucidate the state of Cr interested reader should consult this paper for
in the surface layers of Cr/H-Y zeolite. The binding further details [92].
energy value of the Cr 2p level for chromium in
the zeolite differs from that of bulk chromium 5.5. Catalysis and coke formation
oxide compounds or that of Cr supported on silica.
The counter-Cr 3÷ ions in the surface layers of Y XPS is more and more successfully applied to
zeolite are reduced by heat treatment both in CO problems related to catalysis and coke formation.
and in vacuo to Cr 2+, unlike the Cr 3 + ions in the The quantitative use of XPS intensity ratios,
zeolite bulk, which are not self-reduced. The stable although not yet fully developed, has interesting
252 M. Strcker/Microporous Materials 6 (1996) 235-257

potential applications in the characterization of The splitting is due to the charge donation from
zeolite catalysts having quite complex structural the coke to the framework oxygens of the
features. It is especially noteworthy that, by using FeHNa-Y zeolite. The strongly basic oxygens and
XPS, these features can be examined whatever the the hydrogen-deficient coke are the active center
chemical environment of the metallic elements in the dehydrogenating coked FeHNa-Y zeolite
under consideration, as demonstrated with catalyst. The O ls XPS peak at 527.8 eV for the
FePd/ZSM-5 catalysts prepared by two different coked FeHNa-Y zeolite is the fingerprint peak for
techniques and used for deoxygenation of organic the catalysts active in the dehydrogenation reac-
compounds in the presence of CO [93]. The varia- tions [94].
tions of the (Me/Si)xPs atomic ratios are calculated These investigations have been extended by the
from the peak areas of the corresponding photo- same author to include mordenite and H-ZSM-5
electron transitions by Eq. 7: catalysts. The multiple structure of the Si 2s peak
in the XPS of coked zeolites indicates that the
IM~ OSi Dsi 2Si coke deposit interacts strongly with SiO4 tetrahe-
(Me/Si)xPs = (7)
/Si 0"MeDMe '~Me dra [95].
where I is the XPS intensity, a is the photoelectron The liquid-phase hydration of acrylonitrile to
cross-section, 2 the escape depth and D the effi- acrylamide has been studied using Cu-exchanged
ciency of collection [93]. Y zeolite. Identification of oxidation states reveals
A splitting of the O Is peak in the XPS of a that during dehydration Cu hydroxyl species are
dehydrogenating coked FeHNa-Y zeolite catalyst formed. During the calcination, autoreduction of
has been observed by Kulkarni (see Fig. 11) [94]. Cu 2+ to a lower oxidation state was seen. The
catalyst is reduced almost completely under
hydrogen atmosphere to the Cu ° state. However,
527.8 in the used catalyst, only Cu 2+ in a CuO phase
was observed. During the reaction, it is possible
(b) for the Na in the catalyst sample to migrate from
the surface to pores that are inaccessible to X-ray
radiation during XPS analysis. A comparison of
surface atomic ratios of elements in fresh and used
catalysts indicates that in the used catalyst the
silicon concentration has been reduced consider-
ably, resulting in a lowering of the Si/AI atomic
J ratio. This may also indicate that Cu is preferen-
tially replacing Si on the surface. As the surface

¢
of the used catalyst is rich in Cu, the Cu from the
(a) pores of the zeolite must have migrated to the
surface during the hydration reactions [96].
Rh/Y zeolite catalysts were investigated with
528.4 ,~ XPS, and it was found that a considerable amount
of Rh ÷ was produced as a stable intermediate
during the reduction of Rh 3÷ to rhodium metal
by the heat treatment of the Rh/Y catalysts in
| I
vacuum. It was demonstrated that Rh ÷ in the
520 540 zeolite is active for the hydrogenation and the
Binding energy (eV) ~, dimerization of ethylene and that Rh metal is
Fig. 11. XPS spectra (O ls transitions) of the Fe/Na-Y (a) and
active for the hydrogenation of both ethylene and
Fe/Na-Y (b) (coked) catalysts (reproduced by permission of acetylene. Strong correlations between liquid sys-
Scientific Publishers, Jodhpur, India). tems containing dissolved Rh complexes and solid
M. StScker/Microporous Materials 6 (1996) 235-257 253

Rh/Y zeolite catalysts were found regarding the


active state of Rh and probably the reaction mech-
anism. Rh in the zeolites showed considerable H-offretite ~ 990.2
migration, clustering and sintering during the acti-
vation process. It was suggested that the structure /xl0
of Rh in the zeolite correlates strongly with the
chemical nature of Rh [97].
The hydrogenation of CO has been studied on
a series of Pd catalysts supported on H-ZSM-5,
Na-ZSM-5, H-Y, Na-Y zeolites and SIO2. XPS
results showed no shifts in binding energies indicat-
ing that metallic Pd was the predominant active .,~ I,t~ 987.9
species [98].
The isomerization of n-butane over Y type and g90,3~/xl
offretite zeolites was carried out by St6cker et al.
[99] applying XPS (among others) for the charac-
terization of the zeolites. For Na-offretite, two
types of Na-sites were disclosed through chemical
shifts in the Na ls binding energy and the Na
KLL Auger kinetic energy. After ion-exchange to I I I

H-offretite, only the lower photoelectron binding 996.6 986.6 976.6

energy and the higher Auger kinetic energy sites Auger kineticenergy(eV)
remained occupied, in reduced amounts, though Fig. 12. Na KL2,3L2, 3 Auger electron spectra of H- and
(see Fig. 12). This observation indicates the exis- Na-offretite (AI Kc~ X-rays).
tence of both non-exchangeable and exchangeable
Na-ions in Na-offretite, the former being located
in a more electron-rich environment in the zeolite oxygen atoms become more basic in nature. The
framework [99]. results suggest that the coke was deposited on
Two ZSM-5 zeolites with different Si/A1 ratios SiO4 tetrahedra [ 101 ].
(39 and 74, respectively) were used as catalysts for XPS results concerning the adsorption of
the conversion of methanol. The coke formation organic bases on H-Y zeolites outgassed at
on the external and internal surfaces of the ZSM-5 different temperatures have been reported by
crystals was readily distinguished by using XPS to Defosse and Canesson [102]. The position of the
follow the C/Si ratio. This ratio increased linearly N ls level of aniline adsorbed on H-Y zeolite
at a low rate with internal coke formation, and undergoes a shift of approximately - 2 eV when
exponentially when external coke was formed. the calcination temperature exceeds 400°C, which
Internal coke formation predominated during is the dehydroxylation temperature for the zeolite.
methanol conversion over ZSM-5 until the catalyst The XPS line position seems to depend on whether
was essentially deactivated. External coke was then the adsorption takes place on Bransted or Lewis
formed, probably by the thermal cracking of meth- sites [102].
anol [100]. The reaction of toluene with ethanol
over H-ZSM-5 (Si/AI ratio of about 27) was 5.6. Characterization of acidity and basicity in
followed by XPS analysis of the partially coked zeolites
catalyst. As a result of the coke deposition, the Si
2s and C ls binding energies were decreased by The XPS method has been proposed for the
1.0 eV due to the electron charge transfer from the identification and quantitative determination of
coke to the framework Si- and O-atoms. After the Brensted and Lewis acid sites in zeolites, and
reactivation of the coked sample, the framework inspiring work has been done by Borade et al.
254 M. St6cker/Microporous Materials 6 (1996) 235-257

[103]. The method consists of deconvoluting the Usually such measurements cannot be made by
N ls XPS level of chemisorbed pyridine and IR spectroscopy [103]. Isomorphous substitution
measuring the relative intensities of the peak com- of Fe for Si or A1 in the ZSM-5 framework leads
ponents. It was found that pyridine is chemisorbed to a slight decrease in the binding energy of the N
in three different states on, for example ZSM-5, 1s Lewis component. In the case of boron substitu-
corresponding to N ls binding energies of 398.7, tion in ZSM-5 a decrease in the binding energy of
400.0 and 401.8 eV, respectively. The first peak at all three N ls components was observed [107].
398.7 eV was assigned to the N ls level of pyridine The same research group silylated H-ZSM-5
adsorbed on Lewis sites, while the second and using different agents and investigated the acidity
third were assigned to the N ls levels of pyridine of the samples by XPS of chemisorbed pyridine.
adsorbed on relatively weak and strong Bronsted The results suggest that the removal of a Bransted
acid sites, respectively (an example with Y zeolite acid site produces a framework trigonal A1 atom,
is shown in Fig. 13 [104,105]). Comparisons of which is the Lewis acid site in H-ZSM-5 zeolites.
the concentrations of the various acid sites as Compared with the IR spectra, the XPS spectra
determined from the relative intensities of the N of chemisorbed pyridine are more sensitive in
1s components with IR data showed that XPS has distinguishing Brensted acid sites of different
potential applications in the identification and the strengths. The removal of external silanol groups
quantitative determination of Bronsted and Lewis does influence the distribution of surface acid sites
acid sites in zeolites [106]. Moreover, the XPS by increasing the proportion of strong surface
technique has the ability of providing the relative Bronsted acid sites [108]. The comparison of the
concentrations of strong and weak Bransted sites. ratio of the Bronsted and Lewis acid sites deter-
mined from the relative intensities of the N ls

LL__
peaks with IR spectroscopic data showed that
there is an inhomogeneous distribution of Brensted
and Lewis acid sites in H-ZSM-5 [109].
The large pore zeolites containing 12-membered
ring pore openings designated as omega, beta,
ZSM-20 and mordenite have been investigated for
their acidic properties using N l s XPS of chemi-
sorbed pyridine [110]. Deconvolution of N ls XPS
spectra revealed that all zeolite samples contained
relatively strong Bransted, weak Bronsted and
Lewis acid sites. The binding energy of the N ls
component peak associated with strong Bronsted
acid sites in omega zeolite was found to be signifi-

jk.
cantly higher than those Bronsted acid sites present
in the other zeolites. The results were compared
with normal H-Y zeolite which contained one type
of Brensted acid site and Lewis acid sites. Based
on the N l s binding energy value zeolites are
arranged for their Brensted acid strength as fol-
lows: omega > beta > ZSM-20 > mordenite > H Y
392 397 402 407 412 zeolite [110].
Binding Energy(eV) The acid strength of Mg-exchanged X- and
Y-type zeolites was examined by Viner et al. using
Fig. 13. Deconv01uted N ls XPS spectra of pyridine chemi-
sorbed on H-Y zeolites with different surface Si/A1 ratios of 3.6
XPS. The high activity of these zeolites was
(A), 4.1 (B) and 12.9 (C) (reproduced by permission of Elsevier reflected in their Mg 2p binding energies which
Science B~V., Amsterdam). were higher than those in the oxide and salts. The
M. Strcker/Microporous Materials 6 (1996) 235-257 255

AI 2p and Si 2p binding energies were smaller than alkali exchanged X zeolites [ 113]. One component
those of the oxides [ 111 ]. (400.6 eV) is formed in all spectra (see Fig. 14),
Finally, an XPS method was proposed for the and as it is the dominant peak in the spectrum of
characterization of Lewis basic sites in alkali- Na X zeolite, it is ascribed to pyrrole chemisorbed
cation faujasite zeolites by Huang et al. [112]. The on a surface oxygen adjacent to a sodium cation.
technique consists of deconvoluting the N ls XPS The other major peak (399-401 eV) is assigned to
line of chemisorbed pyrrole and measuring the pyrrole chemisorbed on the basic sites associated
relative intensities of the peak components. It was with the other cation. The binding energy for these
found that, whenever the zeolite sample contains sites is in the order of increasing electropositivity:
two kinds of alkali cations, two different deconvo- Li > Na > K > Rb > Cs. Thus the lower binding
luted N ls peaks corresponding to pyrrole chemi- energy is associated with higher basic strength.
sorbed on these two sites are obtained. Combined Similar results were obtained for a series of alkali
with IR observations, the results offer further exchanged Y zeolites [113].
evidence for the basic sites being the framework
oxygen adjacent to the alkali cations in faujasite
zeolites. The stronger the basic site, the lower the 6. Concluding remarks
N ls binding energy level of the corresponding
peak component is. An N ls peak corresponding It has been shown that XPS is a powerful tool
to polymerized pyrrole species was also detected for the characterization of zeolites and related
in the cases of Li-Y and Na-Y zeolites [112]. An materials. The author is aware that not all contri-
example is demonstrated in Fig. 14 which shows butions published so far could be taken into
the N ls XPS lines of pyrrole chemisorbed on account in connection with the preparation of this
review. However, he hopes that this paper can
provide an overview within this part of surface
characterization of zeolites and related materials
research.

Acknowledgements

The author is indebted to Tordis Whist for


technical assistance in connection with the prepara-
tion of the figures, and the relevant publishers are
gratefully acknowledged for their permission to
reproduce the figures as cited in the text. Financial
support by The Research Council of Norway
(NFR) and SINTEF Chemistry is gratefully
acknowledged.

References

391 396 401 406 411 [1] D. Briggs, Spectrosc. Eur., 5 (1993) 8.
Binding energy, eV [2] J.M. Walls (Editor), Methods of Surface Analysis,
Cambridge University Press, Cambridge, 1989.
Fig. 14. N ls XPS lines of pyrrole chemisorbed on alkali [3] D. Briggs and M.P. Seah (Editors), Practical Surface
exchanged X zeolites (reproduced by permission of Elsevier Analysis, 2nd ed., John Wiley & Sons,
Science B.V., Amsterdam). Chichester/Salle + Sauerl~inder, Aarau, 1994.
256 M. St6cker/Microporous Materials 6 (1996) 235-257

[4] J.W. Niemantsverdriet, Spectroscopy in Catalysis, Verlag [35] B. Herreros, H. He, T.L. Barr and J. Klinowski, J. Phys.
Chernie, Weinheim, 1993. Chem., 98 (1994) 1302.
[5] C.D. Wagner, Faraday Discuss. Chem. Soc., 60 (1975) [36] H. He, T.L. Barr and J. Klinowski, J. Phys. Chem., 98
291. (1994) 8124.
[6] C.D. Wagner, Anal. Chem., 47 (1975) 1201. [37] A.M. Winiecki, S.L. Suib and M.L. Occelli, Langmuir, 4
[7] E.S. Shpiro, G.V. Antoshin and Kh.M. Minachev in D. (1988) 512.
Kall6 and Kh.M. Minachev (Editors), Catalysis by [38] M.J. Remy, M.J. Genet, P.P. Nott6, P.F. Lardinois and
Zeolites, Akad6miai Kiad6, Budapest, 1988, p. 43. G. Poncelet, Microporous Mater., 2 (1993) 7.
[8] T.L. Barr, Prepr., Div. Petr. Chem., Am. Chem. Soc., 23 [39] E. Ponthieu and P. Grange, Zeolites, 12 (1992) 402.
(1978) 82. [40] M. St6cker, J.K. Grepstad and K.P. Lillerud, Catal.
[9] T.L. Barr, Appl. Surf. Sci., 15 (1983) 1. Today, 3 (1988) 97.
[10] T.L. Barr and M.A. Lishka, J. Am. Chem. Soc., 108 [41] F. Delannay and S. Ceckiewicz, Zeolites, 5 (1985) 69.
(1986) 3178. [42] A. Aurottx, H. Dexpert, C. Leclercq and J. Vedrine, Appl.
[11] T.L. Barr, L.M. Chen, M. Mohsenian and M.A. Lishka, Catal., 6 (1983) 95.
J. Am. Chem, Soc., 110 (1988) 7962. [43] E. Alsdorf, M. Feist, H. Fichtner-Schmittler, Th. Gross,
[12] W. GrOnert, M. Muhler, K.-P. Schr6der, J. Sauer and R. H.-J. Jerschkewitz, U. Lohse and B. Parlitz, Ads. Sci.
SchlOgl, J. Phys. Chem., 98 (1994) 10920. Technol., 5 (1988) 127.
[13] T.L. Barr, Zeolites, 10 (1990) 760. [44] A.E. Hughes, K.G. Wilshier, B.A. Sexton and P. Smart,
[14] T.L. Barr, Microporous Mater., 3 (1995) 557. J. Catal., 80 (1983) 221.
[15] J. Stoch, J. Lercher and S. Ceckiewicz, Zeolites, 12 [45] S.L. Suib, G.D. Stucky and R.J. Blattner, J. Catal., 65
(1992) 81. (1980) 174.
[16] Y. Okamoto, M. Ogawa, A. Maezawa and T. Imanaka, [46] J. Dwyer, F.R. Fitch, F. Machado, G. Qin, S.M. Smyth
J. Catal., 112 (1988) 427. and J.C. Vicherman, J. Chem. Soc., Chem. Commun.,
[17] W. G~nert, R. Schl6gl and H.G. Karge, Surf, Interface (1981) 422.
Anal., 20 (1993) 603. [47] R. von Ballmoos and W.M. Meier, Nature, 289 (1981) 78.
[18] V.K. Kaushik, S.G.T. Bhat and D.R. Corbin, Zeolites, 13 [48] E.G. Derouane, J.P. Gilson, Z. Gabelica, C. Monsty-
(1993) 671. Desbuquoit and J. Verbist, J. Catal., 71 (1981) 447.
[19] Th. Gross, U. Lohse, G. Engelhardt, K.-H. Richter and [49] F. Pechar, D. Rykl and L. Mikusik, Zeolites, 2 (1982) 257.
V. Patzelov~i, Zeolites, 4 (1984) 25. [50] M. Narayana, S. Contarini and L. Kevan, J. Catal., 94
[20] B.L. Meyers, T.H. Fleisch and C.L. Marshall, Appl. Surf. (1985) 370.
Sci., 26 (1986) 503. [51] S. Contarini and L. Kevan, J. Phys. Chem., 90 (1986)
[21] J.N. Ness, D.J. Joyner and A.P. Chapple, Zeolites, 9 1630.
(1989) 250. [52] B.A. Sexton, T.D. Smith and J.V. Sanders, J. Electron
[22] V. Andera, L. Kubelkov~, J. NovLkov~i, B. Wichterlov~i Spectrosc. Relat. Phenom., 35 (1985) 27.
and S. Bednfirovfi, Zeolites, 5 (1985) 67. [53] I. Jirka, B. Wichterlov~t and M. Maryska, Stud. Surf. Sci.
[23] I. Grohmann and Th. Gross, J. Electron Spectrosc. Relat. Catal., 69 (1991) 269.
Phenom., 53 (1990) 99. [54] I. Jirka and V. Bos~icek, Zeolites, 11 (1991) 77.
[24] J.-Fr. Tempere, D. Delafosse and J.P. Contour, Am. [55] L.P. Haack and M. Shelef, Am. Chem. Soc. Symp. Ser.,
Chem. Soc. Symp. Ser., 40 (Molecular Sieves-2) (1977) 76. 552 (Environmental Catalysis) (1994) 66.
[25] Q.L. Wang, M. Torrealba, G. Giannetto, M. Guisnet, G. [56] E.S. Shpiro, W. Grllnert, R.W. Joyner and G.N. Baeva,
Perot, M. Cahoreau and J. Caisso, Zeolites, 10 (1990) 703. Catal. Lett., 24 (1994) 159.
[26] J.Z. Shyu, E.T. Skopinski, J.G. Goodwin, jr. and A. [57] S. Xiao and Z. Meng., J. Chem. Soc., Faraday Trans., 90
Sayari, Appl. Surf. Sci., 21 (1985) 297. (1994) 2591.
[27] S. Nakata, K, Nagaoka, S. Asaoka and H. Takahashi, [58] S. Badrinarayanan, R.I. Hegde, I. Balakrishnan, S.B.
Anal. Sci., 2 (1986) 95. Kulkarni and P. Ratnasamy, J. Catal., 71 (1981) 439.
[28] E. Merlen, J. Lynch, M. Bisiaux and F. Raatz, Surf. [59] S. Contarini, J. Michalik, M. Narayana and L. Kevan,
Interface Anal., 16 (1990) 364. J. Phys. Chem., 90 (1986) 4586.
[29] M.J. Remy, M.J. Genet, G. Poncelet, P.F. Lardinois and [60] Kh.M. Minachev, G.V. Antoshin, E.S. Shpiro, F. Vogt,
P.P. Nott6, J. Phys. Chem., 96 (1992) 2614. K.-H. Bayer and H. Bremer, Z. Chem., 17 (1977) 386.
[30] A. Corma, V. Forn6s, O. Pallota, J.M. Cruz and A. [61] F. Steinbach, J. Schtitte, R. Krall, Chr. Minchev, V.
Ayerbe, J. Chem. Soc., Chem. Commun., (1986) 333. Kanazirev and V. Penchev, in D. Olson and A. Bisio
[31] J. K~irger, H. Pfeifer, R. Seidel, B. Staudte and Th. Gross, (Editors), Proc. 6th International Zeolite Conference,
Zeolites, 7 (1987) 282. Butterworth, Guildford, 1984, p. 417.
[32] J. Knecht and G. Stork, Z. Anal. Chem., 283 (1977) 105. [62] P. Mikusik, T. Juska, J. Nov~ikov~i,L. Kubelkov~i and B.
[33] J. Knecht and G. Stork, Z. Anal. Chem., 286 (1977) 44. Wichterlov~i, J. Chem. Soc., Faraday Trans. 1, 77 (1981)
[34] J. Knecht and G. Stork, Z. Anal. Chem., 286 (1977) 47. 1179.
M. St6cker/Microporous Materials 6 (1996) 235-257 257

[63] S.J. Kulkarni, S. Badrinarayan and S.B. Kulkarni, [88] J.C. Vedrine, M. Dufaux, C. Naccache and B. Imelik,
J. Catal., 75 (1982) 425. J. Chem. Soc., Faraday Trans. I, 74 (1978) 440.
[64] A. Yu. Stakheev and W.M.H. Sachtler, J. Chem. Soc., [89] B. Wichterlov~i, L. Krajcikov~i, Z. Tvaruzkovfi and S.
Faraday Trans., 87 (1991) 3703. Beran, J. Chem. Soc., Faraday Trans. I, 80 (1984) 2639.
[65] M. Narayana, J. Michalik, S. Contarini and L. Kevan, [90] A. Corma, M.I. Vazquez, A. Bianconi, A. Clozza,
J. Phys. Chem., 89 (1985) 3895. J. Garcia, O. Pallota and J.M. Cruz, Zeolites, 8 (1988) 464.
[66] J.C. Vedrine, M. Dufaux, C. Naccache and B. Imelik, [91] O.P. Tkachenko, E.S. Shpiro, M. Wark, G. Schulz-Ekloff
Proc. 7th International Vacuum Congr. and 3rd and N.I. Jaeger, J. Chem. Soc., Faraday Trans., 89
International Conf. Solid Surf., 1 (1977) 481. (1993) 3987.
[67] Z. Zsoldos, G. Vass, G. Lu and L. Guczi, Appl. Surf. Sci., [92] A. Sayari, A. Adnot, S. Kaliaguine and J.R. Brown,
78 (1994) 467. J. Electron Spectrosc. Relat. Phenom., 58 (1992) 285.
[68] S.M. Kuznicki and E.M. Eyring, J. Catal., 65 (1980) 227. [93] F. Simard, A. Adnot and S. Kaliaguine, J. Mol. Catal.,
[69] Y.-Y. Liu, W.-J. Zhao, S.-J. Zhang, Y.-Q. Fang and J.-K. 50 (1989) 81.
Hao, J. Nat. Gas Chem., 2 (1993) 88. [94] S.J. Kulkarni, Ind. J. Chem., 29A (1990) 1125.
[70] L.A. Pedersen and J.H. Lundsford, J. Catal., 61 (1980) 39. [95] S.J. Kulkarni, Ind. J. Chem., 29A (1990) 201.
[71] Y.-Y. Liu, W.-J. Zhao, S.-J. Zhang and Y.-Q Fang, Appl. [96] V.K. Kaushik and M. Ravindranathan, Zeolites, 12
Surf. Sci., 59 (1992) 299. (1992) 415.
[72] W. GrOnert, U. Sauerlandt, R. Schlrgl and H.G. Karge, [97] Y. Okamoto, N. Ishida, T. Imanaka and S. Teranishi,
J. Phys. Chem., 97 (1993) 1413. J. Catal., 58 (1979) 82.
[73] T. Arakawa, T. Takata and J. Shiokawa, Chem. Phys. [98] N.C. Saha and E.E. Wolf, Appl. Catal., 13 (1984) 101.
Lett., 72 (1980) 469. [99] M, Strcker, P. Hemmersbach, J.H. R~eder and J.K.
[74] V. Kanazirev, G.L. Price and G. Tyuliev, Zeolites, 12 Grepstad, Appl. Catal., 25 (1986) 223.
(1992) 846. [100] B.A. Sexton, A.E. Hughes and D.M. Bibby, J. Catal.,
[75] I. Grohmann, W. Pilz, G. Walther, H. Kosslick and V.A. 109 (1988) 126.
Tuan, Surf. Interface Anal., 22 (1994) 403. [101] S.J. Kulkarni, H. Hattori and I. Toyoshima, Zeolites, 7
[76] D. Trong On, L. Bonneviot, A. Bittar, A. Sayari and S. (1987) 178.
Kaliaguine, J. Mol. Catal., 74 (1992) 233. [102] C. Defosse and P. Canesson, React. Kinet. Catal. Lett.,
[77] M. Huang, A. Adnot and S. Kaliaguine, J. Am. Chem. 3 (1975) 161.
Soc., 114 (1992) 10005. [103] R. Borade, A. Sayari, A. Adnot and S. Kaliaguine,
[78] M.J. Edgell, R.W. Paynter, S.C. Mugford and J.E. Castle, J. Phys. Chem., 94 (1990) 5989.
Zeolites 10 (1990) 51. [104] R. Borade, A. Adnot and S. Kaliaguine, J. Mol. Catal.,
[79] Kh. Minachev, G.V. Antoshin, Yu.A. Yusifov and E.S. 61 (1990) L7.
Shpiro, React. Kinet. Catal. Lett., 4 (1976) 137. [105] C. Guimon, A. Boreave and G. Pfister-Guillouzo, Surf.
[80] Y.S. Yong, R.F. Howe, A.E. Hughes, H. Jaeger and B.A. Interface Anal., 22 (1994) 407.
Sexton, J. Phys. Chem., 91 (1987) 6331. [106] C. Defosse, B. Delmon and P. Canesson, Am. Chem.
[81] S.L.T. Andersson and R.F. Howe, J. Phys. Chem., 93 Soc. Symp. Ser., 40 (Molecular Sieves-2) (1977) 86.
(1989) 4913. [107] R. Borade, A. Adnot and S. Kaliaguine, J. Catal., 126
[82] E.S. Shpiro, G.V. Antoshin, O.P. Tkachenko, S.V. (1990) 26.
Gudkov, B.V. Romanovsky and Kh.M. Minachev, Stud. [108] M. Huang, A. Adnot and S. Kaliaguine,.J. Chem. Soc.,
Surf. Sci. Catal., 18 (1984) 31. Faraday Trans., 89 (1993) 4231.
[83] J. Strutz, H. Diegruber, N.I. Jaeger and R. Mrseler, [109] F.-S. Xiao, R.-R. Xu and Y.-H. Xu, Chem. Res. Chin.
Zeolites, 3 (1983) 102. Univ., 8 (1992) 337.
[84] E. P~iez-Mozo, N. Gabriunas, F. Lucaccioni, D.D. Acosta, [110] R. Borade and A. Clearfield, Stud. Surf. Sci. Catal., 84
P. Patrono, A. La Ginestra, P. Ruiz and B. Delmon, (1994) 661.
J. Phys. Chem., 97 (1993) 12819. [ 111] H. Viner, H. NoUer and M. Ebel, Z. Phys. Chem. NF,
[85] R. Hoppe, G. Schulz-Ekloff, D. Wrhrle, E.S. Shpiro and 103 (1976) 325.
O.P. Tkachenko, Zeolites, 13 (1993) 222. [112] M. Huang, A. Adnot and S. Kaliaguine, J. Catal., 137
[86] G. Schulz-Ekloff, D. Wright and M. Grunze, Zeolites, 2 (1992) 322.
(1982) 70. [113] R. Borade, M. Huang, A. Adnot, A. Sayari and S.
[87] J.H. Lundsford and D.S. Treybig, J. Catal., 68 (1981) 192. Kaliaguine, Stud. Surf. Sci. Catal., 75 (1993) 1625.

You might also like