Sie sind auf Seite 1von 17

MICROPOROUS

MATERIALS
ELSEVIER

Microporous Materials 11 (1997) 275-291

Nature and strength of acid sites in HY zeolites:


a multitechnical approach
A. BorCave 1,2, A. Auroux

I, C. Guimon 2,*

1 Institut de Recherches sur la Catalyse (CNRS), Avenue A. Einstein, 69626, Villeurbanne, France
UMR (CNRS) 5624, Universite de Pau et des Pays de Iddour, HPlioparc Pau-PyrtnCes, 2 avenue du PrCsident Angot,
64000 Pau, France
Received 30 December

1996; accepted 12 May 1997

Abstract
The combination of several physico-chemical
techniques such as temperature programmed desorption, microcalorimetry, infrared spectroscopy and photoelectron
spectroscopy, together with structural data (by XRD and NMR), has
allowed a detailed study of the acidity of commercial HY faujasites, both non-dealuminated
and dealuminated
by
hydrothermal
treatment. These techniques have been used with different basic probe molecules such as ammonia,
pyridine and 2,6-lutidine (2,6-dimethylpyridine).
They have demonstrated the existence of three different regions of
acid strength. The weak sites associated to NH, adsorption heats between 70 and 130 kJ mol -I are of Lewis and
Brijnsted types, the former being predominent
in the dealuminated
zeolites. The sites of intermediary strength are
essentially the framework bridged hydroxyl groups (heats between 130 and 150 kJ mol -). The strong sites (heats
above 160 kJ mol -l) contain a significant number of extra-framework
Lewis-type sites, but also Briinsted s:tes. The
strength of these sites seems to be due to mutual interactions between Lewis and Bronsted sites. The use of 2,6-lutidine
has furthermore
allowed us to differentiate between two different populations
of Lewis sites. Q 1997 Elsevier
Science B.V.
Keywords:

Zeolites Y; Faujasites; Acidity;

Chemisorption;

Thermodesorption;

1. Introduction
Acid catalysts, and among them zeolites, have
many important applications in the chemical and
petrochemical industries. In particular, dealuminated HY faujasites (ultra-stable Y or USY) are
the most widely used catalysts (FCC processes)
together with ZSM-5 zeolites. It is widely acknowledged that the active centers of these microporous
* Corresponding author. Tel: + 33 5 59843150;
Fax: + 33 5 59923029; e-mail: claude.guimon@univ-pau.fr
0927-6513/97/$17.00 0 1997 Elsevier Science B.V. All rights reserved.
PII SO927-6513(97)00047-3

Microcalorimetry

solids are Bronsted acid sites carried by the zeolite


framework (bridged OH groups) [l-3]. These sites
are present in very high amounts in synthetic Y
faujasites with an atomic Si/Al ratio usual1.ylower
than 3. However, these zeolites do not present a
very good thermal stability due to the high aluminum content, so that it is necessaryto dealuminate
them, either by chemical treatment, or more commonly by hydrothermal treatment (steaming).
These treatments, which partially remove aluminum from the crystal framework (FAI,), thus
improving their stability, decrease the nulmber of

276

A. BorPuve

et al. , Microporous

Materials

acid sites but increase their average strength. It


has been suggested [4,5] that this increase stems
from the preferential loss of the aluminum atoms
which have more than one aluminum as nextnearest neighbor, which are the weakest sites. On
steaming, the aluminum removed from the framework
remains in the cavities (extra-framework
aluminum, or EFAL) as nanoparticles whose surface possesses Lewis acidity [6] related to, among
other species, tri-coordinated
Al atoms or/and
aluminum cations [7]. Furthermore,
the dealumination treatments tend to deteriorate the crystallinity of solids and generate framework defects and
thus additional Lewis sites. These Lewis sites may
have their own catalytic activity [8,9] or may
interact with Briinsted sites, inducing an increase
of their strength and thus a modification of their
catalytic activity [ 10,111.
Therefore, the acidity of zeolites involves relatively complex phenomena, due to this intrinsic
heterogeneity. Its study requires the use of numerous physico-chemical
techniques which, when used
together, can simultaneously provide information
on the number, distribution,
nature and strength
of the acid sites as well as on the way they are
related to the composition, crystallinity and structure of the solid.
The aim of the current work was the detailed
study of the acidity of HY and H-USY zeolites
and the influence of dealumination treatments on
this acidity. For this purpose, besides the techniques allowing the investigation of the structure
(XRD,
NMR, BET), we have used thermal and
Table 1
Characteristics

of studies

I1 (1997)

275-291

spectroscopic techniques in relation with adsorption experiments involving gaseous bases. These
techniques are temperature-programmed
desorption, microcalorimetry,
infrared spectroscopy, and
X-ray
photoelectron
spectroscopy.
The basic
probes selected were ammonia,
pyridine and
2,6-lutidine (2,6-dimethylpyridine),
which differ in
acidity and, more importantly, in size, and for the
latter molecule the accessibility is modified by the
steric hindrance due to the presence of methyl
groups close to the nitrogen atom.

2. Experimental
2.1. Materials
Commercial Y faujasites were supplied by Union
Carbide (LZ-Y62),
Conteka (CBV 712, CBV 720,
CBV 760) and Zeocat (ZF 220, ZF 520).
LZ-Y62, an ammonium exchanged zeolite, was
calcined at 573 K in a dry air stream, then activated
at 673 K in the He flow prior to adso:rption (like
all the other zeolites). The calcined sample was
named LZ-Y62c.
The CBV 100 set (712, 720, 760) corresponds
to ultra-stabilized
Y zeolites (USY)
obtained by
steaming.
ZF 220 and ZF 520 are also H-USY. The sample
ZF 220 has a large quantity of extra-framework
aluminum (EFAL)
species (overall Si/Al atomic
ratio: 2.7; framework
Si/Al ratio: 20) which have

zeohtes

Zeolite

LZY62r

LZ-Y62C

CBV

Overall SiiA13
Surface Si/A14
Sub-surface
Si/Als
Crystallinity6
(%)
BET (N2) surface (mZ g-)

2.6
2.8
2.8
100
948

2.6
2.8
2.7
76

5.8
3.5
5.1
33
700

NH,Y
form.
Calcined
(dry air) at 573 K.
3Elemental
analysis.
4XPS analysis of as-receivzed samples.
5XPS analysis of ground samples.
6XRD data by reference to the best crystallized

sample

(LZ-Y62)

712

CBV
12.8
26
19
24
730

720

CBV760

ZF 220

ZF 520

24
19
26
19
720

2.7
1.0
2.7
36
700

20
20
20
40
800

A. Bereave

et al. / Microporous

been removed by acid leaching in the case of ZF


520 (overall and framework
Si/Al ratios both 20).
The relevant data on these zeolites are summarized in Table 1.

2.2. Temperature-programmed

desorption

All activations, adsorptions and thermodesorpunder


helium
Ilow
tions
were
performed
(25 cm3 min- ) and were monitored with a conductivity cell.
After activation at 673 K, adsorption was carried out in the pulse mode at 393 K (or other
temperatures
in the case of LZ-Y62c;
see text):
successive small doses (0.4 cm3) of gas were sent
onto the sample (between 50 and 100 mg) up to
saturation. After the sample was swept with pure
helium for 1 h at 393 K and cooled to room
temperature, temperature-programmed
desorption
(TPD) was performed between 303 and 673 K
(4 K min-I) and the desorbed ammonia was measured by back titration (with HCl and NaOH).

2.3. Calorimetry
The heats of adsorption
were measured in a
heat flow microcalorimeter
of the Tian-Calvet
type (C80 from Setaram), linked to a volumetric
line allowing the introduction
of small doses of
ammonia or pyridine. The equilibrium pressure
after each introduction
of gas was measured by
means of a differential
pressure gauge from
Datametrics.
Successive doses were sent onto the
sample until a final equilibrium pressure of 67 Pa
was obtained. Before adsorption, the samples were
evacuated at 673 K overnight and the adsorption
temperature was maintained at 393 K in order to
limit physisorption.
The amount of strong sites
associated with heats above 120 kJ mol was evaluated from the difference between the primary
isotherm and the secondary isotherm obtained
after
desorption
under
secondary
vacuum
( 10 m6 mbar) at 393 K and readsorption
of the
base. This difference is named irreversibly chemisorbed amount (I&).

Materials

I1 11997)

275-291

2.4. Fourier transformed

271

infrared spectroscquy

Self supporting
wafers
(7 mg cm-)
were
calcined overnight under 0, in the IR cell at 673 K
and then evacuated under vacuum (2 h). After
cooling to room temperature and adsorption of
pyridine, the cell was evacuated for 1 h at several
temperatures (298, 373, 473, 573 and 673 K). In
each stage the infrared spectrum was recorded
at room temperature
with a Bruker
IFS 48
spectrometer.
2.5. Photoelectron

spectroscopy

After adsorption of a base (ammonia, pyridine,


lutidine with the TPD device). photoelectron
spectroscopy
(XPS) analyses were performed at
room temperature on samples ground in an agate
mortar [ 121 with a SSI 301 spectrometer using a
monochromatic
and focused (spot diameter of
600 urn, 100 W) Al Ka radiation
(1486.6 eV)
under a residual pressure of 5 x lo-* Pa. Charge
effects were compensated by the use of a flood gun
(5 eV). The hemispherical analyzer functmned in
constant pass energy of 50 eV for high resolution
spectra and 150 eV for quantitative analyses. The
experimental bands were fitted with theoretical
bands (80% gaussian, 20% lorentzian) with a least
squares algorithm using a non-linear baseline [ 131.
Quantitative
analyses were performed using the
appropriate Scofield factors [ 141. Binding energies
(BE) were determined by reference to the Si 2p
band (Table 2) in agreement with literature. In all
spectra, the widths (FWHM)
of the different peaks
were remarkably constant (Table 2). Hence a constant FWHM of 1.9 eV was adopted for the theoretical components in the decomposition
of the
Nls band.

3. Results
3.1. Composition and crystallinity

of the zeolites

The thermal treatments affect dramatic:ally the


crystallinity of the zeolites (Table 1). While steaming has a very marked effect, even a mild calcination (573 K) under dry air decreases the cristallin-

278

A. Borhzve

Table 2
Binding energies
Zeolite

Si 2p
Al 2p
Na 1s
0 IS

and widths

BE (eV)(+O.I

et al. i Microporous

of XPS peaks associated

with

Materials

the different

II (1997)

elements

275-291

of zeolites

studied

eV)

LZ-Y62c

CBV

103.0
14.9
1072.3
532.2

103.1
74.9
1072.3
532.2

FWHM
712

CBV

720

CBV

LO3.3
14.9
1072.5
532.5

ity of the parent zeolite (LZ-Y62). This low crystallinity of LZ-Y62c is probably due to contact
with Hz0 after calcination and results from a
partial collapse of the framework. Similarly, the
specific surface area tends to decrease with dealumination due to the presence of extra-framework
fragments in the pores and channels, as well as the
possible formation of mesopores. The presence of
these fragments and the partial amorphization
of
the dealuminated samples (particularly CBV 760)
are also confirmed by the 27A1 NMR spectra
(Fig. 1). Prior to measuring these spectra, the
samples were rehydrated since the dehydrated zeolites contain NMR invisible aluminum [ 151. The
spectra present a main peak around 56 ppm corresponding to tetrahedral framework aluminum, and
a peak near 0 ppm which can be attributed to
octahedral aluminum characteristic of EFAL [ 161.
The proportions of these two types of aluminum
are reported in Table 3. A peak of low intensity
can also be observed around 30 ppm, but it is
difficult to quantify due to the proximity of the
spinning side bands. This band has already been
observed on an LZ-Y62 sample after steaming by
Yang [ 171, who found it to decrease after washing
by HCl and thus interpreted it as related to the
presence of extra-framework entities. Corma et al.
[ 181 and Fripiat et al. [ 191 have proposed that this
band may be due to pentacoordinated aluminum
or to aluminum in a distorted tetrahedral configuration. Klinowski et al. [20] and Freude et al.
[21,22] have assigned this line to non-framework
AlOOH with tetra-coordinated aluminum close to
framework oxygens.
Table 1 gives the Si/Al atomic ratios measured
by XPS for the raw samples and for manually
milled samples. The first values thus correspond

103.5
74.9
1072.8
532.8

760

ZF 220

ZF 520

103.1
14.6
1072.6
532.0

103.4
74.8

120

Fig. 1. *Al

SO

NMR

40

spectra

(+O.l

eV)

1.8
1.8
1.8
2.2

532.6

2.G

(eV)

40-T

of the zeolites

studied.

to the actual surface of the crystallites (the depth


of XPS analysis is of the order of 5 nm), while the
second values correspond to an intermediary subsurface between the surface and the bulk [ 121. The
zeolite with a high Al content (LZ-Y62) and the
dealuminated and washed sample (ZF 520) appear
to be homogeneous between the surface and the
bulk, whereas the other samples display heterogeneities. For instance, CBV 712, CBV 760 and ZF

A. Bortave

et al. / Microporous

coordinated)

to octahedral

Table 3
Proportions
of tetrahedral
spectra (Fig. 1)

(tetra

Zeolite

LZ-Y62c

CBV

13

15

Octahedral

Al (%)

712

Materials

II (1997)

(hexacoordinated)

CBV

275-291

extra-framework

720

CBV

38

25

220 present a higher concentration


of aluminum
at the surface. This means that the EFAL probably
tends to migrate towards the surface during the
dealumination treatment. On the contrary, CBV
720, in spite of its high content of hexacoordinated
aluminum (Fig. I), is visibly more thoroughly
dealuminated at its surface than in the bulk.

760

219

aluminum

from

AI

NMR

ZF 220

ZF 520

10

443

3.2. TPD results


This technique provides information
on the
number and strength distribution
of acid sites.
However it must be noticed that the data on the
strength distribution
may be affected by certain
diffusion problems. Meanwhile,
the number of
sites only takes into account the sites which can
be desorbed below the maximum desorption temperature, viz. 673 K in our case. So the sites which
are still covered above that temperature, i.e. the
strongest sites, are not accounted for. Finally,
though some attempts have been made to improve
the selectivity of TPD [23], this technique provides
only little information on the nature of sites.
3.2.1. LZ- Y62 zeolite
Fig. 2 shows the TPD curves of ammonia
adsorbed on the non-dealuminated
zeolite LZ-Y62
before (LZ-Y62)
and after (LZ-Y62c)
calcination.
The calcination of this sample not only decreases
its crystallinity (Table 1) but also affects its acidity.
Globally, TPD of NH, reveals a lower number of
sites (2.2 mmol g-l). Furthermore,
the first peak
is shifted towards the lower temperatures (maximum at 443 K instead of 480 K), which is indicative of a decrease in the average strength of the
weak sites. As for the dealuminated samples (vide
infra), this decrease of acid strength is related to
the formation
of amorphous
or, at least, less
crystalline zones. The second band is not shifted,
however it tends to widen towards high temper-

LZY62

LC-Y

62

303

393

673

Fig. 2. TPD of non-dealuminated


uncalcined
(LZ-Y62)
and
calcined (LZ-Y62c)
HY after adsorption
of NH, at 333 K.

atures, which can be explained by the appearance


of strong sites. Thus, it seems that calcination
mainly generates sites with a particularly
weak
acidic character
and sites with an increased
strength. Some of these could not to be detected
because of the desorption
temperature
range
(T,,,<673 K).
A method among others for quantifying
the
strength distribution
of the sites consists in performing a series of adsorptions at various temperatures, each followed by a TPD. We have thus
performed adsorptions at 393, 423, 473, 5.23, 573
and 623 K on sample LZ-Y62c. The corresponding
TPD curves are represented in Fig. 3, and the
amount of NH, desorbed for each curve is reported
in Table 4.
In order to visualize more clearly the various

280

A. Bortave

et al. / Microporous

Materials

II ( 1997) 275-291

443
038

17

2 0.6
2
5 0,4

570
505
580

590

f--l

0
423

473

523

513

T plateau

B
C ZP

Fig. 4. Amounts
tion temperature

623

673

(K)

of desorbed NH, as a function


according
to the TPD reported

01 the adsorpin Fig. 3.

460

303

393

723

Fig. 3. TPD of LZ-Y62c


423 K (B). 473 K (C),

-I- (K)

after adsorption
of NH, at 393 K (A),
523 K (D), 573 K (E) and 623 K (F).

families of sites, it is convenient to present the


data as a bar graph (Fig. 4) where each bar
represents the difference between the amounts of
NH, desorbed during two consecutive TPDs ( A-B2
B-C,. . .). Indeed, these values correspond to the
number of sites having released NH, in a given
50 K temperature range (between 393 and 423 K
for the first bar, 423 and 473 K for the second
one,. . .). The two populations of sites identified by
the first TPD (Fig. 2), namely weak (-423 K)
and medium strength sites (573 K), are clearly
evidenced by the bar graph.

580

CBV760

>

393

3.2.2. Dealuminated zeolites


The TPD curves for ammonia on the dealuminated zeolites CBV 712, CBV 720 and CBV 760.
plotted in Fig. 5, also show the existence of two
Table 4
NH,
amounts
desorbed
during
adsorption
at various temperatures

Tads( K 1
y (mm01 g)

the TPD performed


(T,,,) for LZY62c

after

393
2.2

423
1.4

413
I.1

523
0.9

573
0.3

623
0.2

Fig. 5. TPD

673
(NH,)

of the dealuminated

zeolites

T(K)
CBV.

well-separated site populations of different


strengths. The first one is located in the band
between 415 and 460 K, which is a lower temperature than for the non-dealuminated uncalcined
LZY62 zeolite (Fig. 2). Thus, it appears that the
average strength of the weak sitestends to decrease
when the aluminum content and the integrity of
the zeolitic structure decrease. This weakening,
which can be correlated with the enrichment of

A. Bereave

et al. , Microporous

Materials

the sample in extra-framework


species, had already
been reported in previous studies [ 24,251.
The shift of this band towards lower temperatures has been explained in the literature [26,27]
by the readsorption
of NH, molecules on neighboring sites, a phenomenon which becomes more
important as the site concentration is higher. It is,
however, difficult to ascribe the observed shift
entirely to this phenomenon, since the second band
does not exhibit a similar evolution. Indeed, the
second family of medium strength sites corresponds to desorption temperatures that are slightly
higher than for LZ-Y62 (between 580 and 605 K
instead of 570 K). On the contrary, we find the
same evolution between the dealuminated zeolites:
Tmax decreases with the degree of dealumination.
This indicates the complexity of the desorption
phenomena for these samples. As expected,the
proportion
of these sites increases with the $/Al
ratio, whereas the total population
decreases
(Table 5).
The very weak acidic character of the aluminumcontaining debris generated by the dealumination
treatments is well evidenced by the TPD curves of
the two samples ZF 220 and ZF 520 (Fig. 6),
whose only difference lies in the presence of these
debris in the first sample and in their near absence
in the second. Indeed, the relative number of sites
uncovered at low temperatures
(450-470 K) is
much higher for the ZF 220 zeolite than for ZF
520: the ratio of the number of weak sites to the
number of medium strength sites is around 2 to 1
for the first, and 0.5 to 1 for the second sample.
This increase in relative concentration is confirmed
by the titration of ammonia desorbed up to 673 K:
the total amounts of desorbed ammonia have
similar values (0.65 and 0.5 mmol g-l for ZF220
and ZF520, respectively) though the first one contains 7.5 times more aluminum than the other one.

Table 5
TPD (NH,)

data for the zeolites

275-291

2x1

393

673

Fig. 6. TPD

(NH,)

of the zeolites

To

ZF.

This fact indicates that most of the EFAL brings


about a weak acidity. In view of the two curves of
Fig. 6, it seems furthermore
that the apparent
concentration
of sites of intermediate strength
(desorption
between 523 and 673 K) is slightly
lower for ZF 220 than for ZF 520. This could be
explained by a lower accessibility of the sites of
the first zeolite due to the presence of aluminumcontaining debris in its cavities.
3.3. Microcalorimetry
Calorimetry
is a very powerful
technique in
catalysis, as it allows the elaboration of sc:ales of
active site strength using direct measurements of
the heats of reaction between the catalyst and a
probe molecule [28-311. Furthermore,
when coupled with volumetric measurements, it gives; access
to the concentration
of sites on the surface of the
solid. As with TPD, the involved energies of
reaction depend both on the acidic solid and the
basic probe molecule, so that it seemed interesting

studied

Zeolite

LZ-Y62

CBV

T, (K)

480
570
3.4

460
605
1.1

T, (K)
odes (mm01 g - 1

-I

303

II (19973

712

CBV
450
600
0.4

720

CBV 760

ZF 220

ZF 520

415
580
0.2

450
590
0.65

470
590
0.5

A. BorPave

282

et ai. : Microporous

to choose again two bases that differ in size


and basicity,
namely ammonia and pyridine.
Fig. 7Fig. 8 illustrate
the variations
of the
differential heats of adsorption of these two bases
on the four zeolites LZ-Y62c, CBV 712, CBV 720
and CBV 760 as a function of coverage (adsorbed
amount).
Generally
the uncertainty
on the
differential heats is + 4 kJ mol - [ 321.
3.3.1. NH, adsorption
Although
the precision of the measurements
of the initial heats of adsorption
is mediocre,
it appears that the strongest sites correspond
to particularly
high heats (between
180 and
230 kJ mol-i).
With
the exception
of the
most strongly dealuminated zeolite (CBV 760),
the curves
present
a near plateau around
140 kJ mol -I, usually ascribed to adsorption on a
homogeneous
population
of Bronsted
sites of
medium strength
[29,33-361.
Afterwards
the

0-i
'-.L7332

20

-Cav712

cBv120

-cw760

----__-__I-__I-----0

A,Hwr

500

Fig. 7. Differential
studied zeolites.

3oo Q lkJ/moll
I
+--LZY62
2% r,

heats

of adsorption

-i

-c&712

CBV720

of ammonia

cBv760

200
150
-y-m--'--m--

. ...'-.
-.

50L---------a-----------j&H,,,&
0'
0

WI

I530

I
2000

v ll!lKhJl

Fig, 8. Differential
studied zeolites.

heats

of adsorption

of pyridine

I1 / 1997) 2 75-291

differential heats decrease to 70 kJ mol -I, a value


often considered as the limit between chemisorption and physisorption
in the case of ammonia.
These curves thus evidence the existence of three
populations of sites with different strengths: the
weak sites (Q below 120-140 kJ mol -), the sites
of medium strength (Q between 120.-140 and
150- 160 kJ mol - ), and the strong and very strong
sites (Q above 150-160 kJ mol - ). These values
are only approximate, mostly in the case of dealuminated zeolites, due to the relatively steep slopes
of the corresponding curves. It is nonetheless possible to evaluate roughly the importance of each
of these populations (Table 6). It should be noted
that these evaluations are well in agree:ment with
the TPD results, considering that the latter only
account for weak and intermediate sites, since the
strongest sites still retain NH, at 673 K ((the maximum desorption temperature).
In agreement with the TPD and NMR curves,
the relative site populations of the two samples
LZ-Y62c and CBV 712 are quite close., although
the global concentrations are not of the same order
of magnitude (the first sample contains more than
twice the number of sites of the second). By
contrast, the two most dealuminated zeolites present fewer weak sites but more very strong sites.
Furthermore it appears that CBV 760 contains the
lowest amount of intermediate sites, which is consistent with its apparent amorphizatison
if one
considers
that these sites are primarily
of
Briinsted type.

on the

100

Materials

on the

3.3.2. Pyridine adsorption


As previously shown in the case of other zeolites
(ZSM-5, mordenite) [37], it is difficult to compare
ammonia and pyridine adsorptions,
because of
their difference in basicity and the specific interactions of each of these molecules with the host
zeolite. The adsorption heats must indeed be correlated not to the basicity of the probe molecules in
liquid phase or in aqueous solution (where NH, is
more basic than pyridine by about 20 kJ mol - ,
being 9.3 and 5.2 pKa, repectively [3X,39]) but to
basicity in the gas phase, which can be correlated
to the proton affinities (PA). Indeed, the PA value
of pyridine is 922.2 kJ mol -l, while that: of ammonia is only 857.7 kJ mol- [40].

A. Bortave
Table 6
Evaluation
(%) of the population
temperatures
(K) of the maxima
ammonia
and pyridine
Acid

strength

Weak
Intermediate
Strong
Krr (NH,)
V,, (pyridine)

et al. / Microporous

Materials

I1 (1997)

275-291

283

(in terms of strength)


of the acid sites from the differential
heat curves (Fig. 7) and corresponding
of NH, desorption
in TPD experiments
(Figs. 2 and 5); irreversible
adsorption
K:,, (mmol g-l) of

LZ-Y62c

CBV

45
39
16
1.9
1.3

50
38
12
0.5
0.6

712

The other parameters which have to be taken


into account are their size (the kinetic diameters
are 0.533 and 0.375 nm, respectively),
and the
secondary interactions between these adsorbates
and the zeolite structure (hydrogen bonds may
lead to the formation
of a monodentate
with
pyridine and a bidentate with ammonia), as shown
by Parrillo et al. [37,41].
These differences are reflected by the adsorption
heats which are between 20 and 30 kJ mol -r higher
in the case of pyridine than in the case of ammonia
[42], including the chemisorption
limit. For the
three dealuminated zeolites (CBV), the results of
pyridine adsorption are close to those for ammonia. However, for LZ-Y62c there appears to be a
deficiency in pyridine chemisorption,
which may
be due to the size of this molecule with respect to
the concentration
of acid sites. Furthermore,
the
adsorption
heats of the sites of intermediate
strength tend to decrease with coverage (contrary
to NH,). An explanation could be the formation
of H-bonded species (pyridine with hydroxyl group
of LF type) which would
weaken the acid
strength of (HF) OH groups [43].
As shown already with NH, adsorption and in
agreement with the literature [44], these TPD
curves
indicate
three regions,
heats above
180 kJ mol-
often assigned
to Lewis
sites
[33,45,46], heats in the range 130-180 kJmol-
commonly associated with Briinsted sites [47-501
and heats between 100 and 130 kJ mol-
attributed in the literature to weak Lewis acidity for
silica-supported
oxides [ 491. The heats lower than
100 kJ mol - are due to hydrogen-bonded
pyridine ( physisorption)
[48-501.

CBV
25
50
25
0.2
0.4

720

CBV
33
33
33
0.1
0.1

760

Tde, (TPD)
4155460
570.-605
3,673

3.3.3. Volumetry
As said above, the difference between th(e primary and secondary adsorption isotherms provides
an evaluation of irreversible adsorption (I{,,) in
the chosen operating conditions (393 K, 10d4 Pa
between the two measurements),
which approximatively corresponds to the adsorption associated
with heats higher than 120 kJ mol - .
The measured values of Q, (Table 6) are: compatible with the previous observations. While the
F,, values for pyridine adsorption on the dealuminated zeolites are equal to or larger than those
for NH,, the order is reversed with LZ-Y62c. In
the first case, the sites are relatively isolated so
that the basicity factor is more pronounced than
the steric factor, while in the second case the site
density is much higher, so that the steric factor
becomes more important.
This is in full in
agreement
with
the observations
madle by
Makarova et al. [36] using FTIR, namely that the
number of Briinsted sites titrated by pyridine is
constant for all zeolites presenting Si/Al atomic
ratios below 4.
3.4. Infrared spectroscop.v
Beside NMR, infrared spectroscopy
is by far
the most commonly used spectroscopy
technique
for the study of the acidity or basicity of catalysts
[51-531. Most studies of surface acidity rely on
the use of basic probe molecules such as carbon
monoxide [ 54,551, or most frequently pyridine [l3,56,57]. Ammonia has been employed in qualitative studies [58,59], but is less interesting due to
the presence of certain overlapping bands [60].

284

A. Borhwe

et al. / Microporous

The band characteristic of the hydroxyl groups


at 3640 cm- (OH located in the large cavities) is
very sensitive to pyridine adsorption on LZ-Y62c
(Fig. 9). At 673 K its intensity is not completely
restored, which shows the strength of these sites.
By contrast, the LF band at 3550 cm-- associated
to the hydroxyl groups located in the hexagonal
prisms and the sodalite cagesis much less affected
by pyridine adsorption, which is in agreement with
the size of this molecule and the previous results.
Finally, the band at 3745 cm- attributed to the
very weakly acidic silanol groups [61] is affected
by pyridine adsorption. This has been ascribed to
a perturbation of these groups by the pyridine
molecules adsorbed on neighboring Lewis sites.
As for the region of OH vibrations (Fig. 9), we
give in Fig. 10 an example of the evolution of the
bands characteristic of vibrations of the pyridine
ring after adsorption and desorption at various
temperatures. The significant physisorption that
occurs at room temperature (band at 1445 cm-)
vanishes completely above 373 K. Note the pres-

Materials

II (1997)

275-291

298 K
373K
473 K
573K
673K
i 9aa

: 30a

i7a0

!6aa
YRVENUtlSER

15aa

,488

4
I3aa

CM-,

Fig. 10. IR spectra of bending and ring-stretching


in LZ-Y62c
saturated
by pyridine
at room temperature
(298 K) followed
by evacuation
at increasing temperature.

ence of a shoulder (high frequency side) only


visible above 573 K. This shoulder has already
been noticed in the case of Y faujasites [60] and
mazzites [62], and has been attributed to the
formation at high temperature of iminium ions
generated by nucleophilic attack of the framework
oxygen followed by protonation of the pyridine
molecules adsorbed on Lewis sites. A.bove 473 K
we can observe a weak band at 1440cm- which
disappears at 673 K. This band has been assigned
by Khabtou et al. [43] to pyridine H-bonded to
LF hydroxyl groups. As the desorption temperature increases,the balance between the intensities
of the two bands at 1545cm- (BriSnsted) and
1455cm- (Lewis) changes in favor of the latter.
It is to be noticed that these bands are still present
at 673 K, which confirms the existence of strong
Bronsted and Lewis sites. This evolution is
common to all the samplesstudied. We will come
back to this point during the discussion.
3.5. X-ray photoelectron spectroscopy
Fig. 9. IR spectra of hydroxyl
groups in LZ-Y62c
then saturated
by pyridine
at room temperature
lowed by evacuation
at increasing temperatures.

(reference),
(298 K) fol-

The potential of XPS in the study of the surface


of materials in general and catalysts in particular

A. Bortave

et al. / Microporous

Materials

is well recognized. Besides the classical determination of the superficial chemical composition
(see
Table 1) and in the manner of spectroscopies like
FTIR, this technique can be applied to studies of
the surface reactivity of solids by using the adsorption of probe molecules. This method proposed 20
years ago by Defosse and Canesson [63] has been
used by various authors in the study of various
families of zeolites, including HY [ 12,641. Like
FTIR, XPS allows the determination of the nature
of the sites (Bransted
or Lewis) and of their
relative concentrations.
In the case of a nitrogencontaining basic probe (NH,, pyridine, lutidine),
the binding energy (BE) associated to the nitrogen
atom (Nls) is indeed a function of its neighborhood. When the base reacts with a Briinsted site
to form an ion (ammonium, pyridinium or lutidinium), the value of BE (Nls) is located between
402 and 402.8 eV. When it forms a complex with
a Lewis site, BE is between 399 and 401.5 eV
depending on the chosen base and the induced
charge transfer. Given the natural width of XPS
bands, the experimental signal corresponds
to a
more or less asymmetrical
band that needs to
undergo deconvolution.
In a previous study [ 121
we had presented the XPS (Nl s) spectra of some
HY zeolites (LZ-Y62,
CBV 712 and CBV 760)
after NH, adsorption at various temperatures. It
had then been noticed that, through monitoring
of the atomic N/Al ratio, the XPS results on milled
samples were in good agreement with the TPD
results regarding the relative proportions
of weak
and strong sites, and that at 673 K between 13
and 22% of the acid sites remained covered by
NH,. Here we compare the XPS results collected
Table I
XPS data after adsorption
of the bases at 393 K. N/Al
sites covered by the basic molecules (L + B = 100%)

275-291

285

after adsorption
at 393 K of the three probes
(ammonia, pyridine and 2,6-lutidine) (Table 7). In
this table we have reported the atomic ratios N/Al
associated to the proportion of titrated sites among
the total aluminum content and N/( Si + Al) which
is related to the concentration
of the base in the
surface and subsurface of zeolites, as well as the
proportion
of base adsorbed
on Lewis sites.
deduced from the spectra of which we give an
example in Fig. 11.
As already observed during the microcalorimetric study, the size of the probe molecule considerably affects its adsorption in the case where
the site concentration is high (Si/Al < 6), while its
effect is much less significant with the most dealuminated zeolites. Indeed pyridine, like lutidine, is
adsorbed twice less than ammonia on LZ-Y62,
and N/Al changes from 0.65 to 0.3. This last value
is in agreement with those measured by Borade
et al. [64] under similar conditions. If the ratio
associated to pyridine on the CBV 712 sample
remains lower than that associated to NH, (0.48
instead of 0.66), nearly no difference is observed
for the other more dealuminated zeolites.
In agreement with the FTIR spectra and a NMR
study by Shertukde et al. [65], all zeolites: even
those with the highest crystallinity (LZ-Y62),
present a notable proportion
of Lewis sites. The
evolution of the number of these sites titrated by
the three bases is interesting, and evidences the
existence of two groups with different beharviors.
For the first one (LZ-Y62, CBV 720 and ZF 520),
the number of Lewis sites titrated by pyridine is
proportionally
lower than that of sites titrated by
NH,, while the opposite behavior is exhibited by

(T = Si + Al) are atomic

ratios.

Pyridine

NH,

LZ-Y62
CBV 712
CBV 720
CBV 760
ZF 220
ZF 520

and N/T

II (1997)

L% are the proportions

of the Lewis

2,6-lutidine

NiAl

N/T

L%

N/AI

NIT

L%

N/Al

N/T

L %

0.65
0.66
0.57
0.68
0.12
0.57

0.16
0.09
0.03
0.02
0.05
0.03

25
35
35
30
35
30

0.3
0.48
0.56
0.59
0.08
0.55

0.08
0.07
0.03
0.02
0.03
0.03

20
25
25
35
40
20

0.3
0.39
0.51
0.53
0.11
0.56

0.08
0.06
0.02
0.02
0.04
0.03

5
5
10
25
25
10

A. Boriave

286

et al. / Microporous

Materials

11 (19971275291

LZ-Y62
Fig. 11. Nls

XP spectra

after

adsorption

CBV760
(393 K) of pyridine

the two samples CBV 760 and ZF 220 which is


related to the accessibility of these sites and the
size of probe molecules.
Switching over to lutidine adsorption, the same
evolution is even more visible. Recall that, due to
the steric hindrance generated by the two methyl
groups neighboring the nitrogen atom, this base is
expected to react much more easily with protons
(BrGnsted sites) than with aluminum atoms (Lewis
sites). This specificity was evidenced by FTIR [66]
on two HY zeolites partially dealuminated by
isomorphous
substitution
or by hydrothermal
treatment (followed
by acid leaching): on the
second sample (with a composition close to that
of CBV 712) the extra-framework
Lewis sites are
still accessible to pyridine after previous adsorption
of 2,6 lutidine and formation of lutidinium species.
This is in agreement with the observed behavior
of the first group of our samples, as evidenced by
the Nls spectra of LZ-Y62 (Fig. 11). However,
this is much less true for the second group
( XP-spectrum of CBV 760 in Fig. 11 ), even if the
number of Lewis sites titrated by lutidine is always
lower than that titrated by pyridine or NH,. Thus,
it seems that the use of these bases with XPS
accounts for the existence of two types of Lewis
site, one accessible
to molecules
such as

and 2,6-lutidine

on LZ-Y62

and CBV

760.

2,6-lutidine, the other accessible only to NH, and


to a lesser extent to pyridine.
For the first group of samples, while the number
of Lewis
sites titrated
after adsorption
of
2,6-lutidine is low, the total number of titrated
sites (Lewis + Briinsted) is nearly the same as with
pyridine. Thus, it seems that the number of
Briinsted sites titrated by lutidine is higher than
that with pyridine. This may be explained in two
ways. On the one hand, 2,6-lutidine is a stronger
base than pyridine and is thus able to titrate
weaker sites which in ultra-high vacuum conditions
do not retain pyridine. On the other hIand, especially in zeolites with high aluminum content whose
site density is thus high, Lewis and Briinsted sites
can be very close to each other, so if a strong
Lewis site is complexed by a large molecule like
pyridine, the weaker neighboring BrMsted site is
no longer accessible to a second molecule. Since
2,6-lutidine does not interact with 1:hese strong
Lewis sites, a higher proportion of Brijnsted sites
is thus accessible.
4. Discussion
The comparison of the data obtained from the
various techniques used here allows one to obtain

A. BorPave

et al. / Microporous

a more complete picture of the acidity of HY and


H-USY zeolites, including in particular the number
of acid sites, their strength and their nature, as a
function of the dealumination treatments.
Acid sites can be subdivided into three families
of different strengths.
Indeed, ammonia TPD
reveals the presence of weak sites (desorbing
ammonia below 523 K) and sites of medium
strength (between 523 and 673 K), while FTIR
and XPS spectroscopies
show the presence of
strong sites, still covered by NH, above 673 K.
Heat flow microcalorimetry
(NH,) has allowed a
quantification
of the strength
of these sites
(between 70 and 130 kJ mol- for the first ones,
then between 130 and 150 kJ mol - , and finally
above 150 kJ mol- ). The distribution of the sites
according to their nature in each of these populations can be evaluated by both FTIR and XPS.
Indeed, the ratio of the number of Briinsted
sites to that of Lewis sites (B/L) as a function of
jesorption temperature can be estimated using the
FTIR spectra, based upon the intensities of the
bands at 1545 and 1455 cm-. The values of the
B/L ratio and their evolution represented in Fig. 12
must only be considered as estimates, because of
the uncertainty in the extinction coefficients. For
our computations we have chosen E,/E, = 1.15 [ 641,
but several other values have been proposed in the
literature [43,67]. It appears that between 373 and
473 K desorption occurs predominantly
on Lewis
sites (B/L increases), while above this temperature
and up to 673 K, the desorbed ammonia is mostly
associated with Bronsted sites (B,/L decreases). At

523

423
Desorption

623

523
tempcmture

(K)

Fig. 12. Briinsted


sites/Lewis
sites ratios as a function
of the
evacuation
temperature
measured by FTIR (E&~ = I. 15).

Materials

II f 1997) 275-291

287

this latter temperature there remains a noticeable


proportion of Lewis sites.
It is possible to study by XPS the NH, adsorption mode as a function of the coverage, i.e. of the
amount of base sent onto the zeolite. As previously
we discuss the results obtained on sample LZY62;
the other zeolites give rise to qualitatively similar
results.
Fig. 13 displays the concentration
of adsorbed
NH, (from the N/Al ratio) as a function of the
volume sent over the zeolite, as determined by
XPS, and, for comparison, as determined by TPD.
For this latter technique, the N/Al ratio includes
not only the amount of NH, desorbed up to
673 K, but also the amount remaining adsorbed
at that temperature, evaluated by XPS (at 673 K,
N/A1=0.13).
After the first dose (0.35mmolg-l),
the two
techniques yield very similar results (N/A.1 between 0.2 and 0.25), which is compatible with an
adsorption on the strongest sites. At higher coverages, the two curves become more and more
widely
separated,
until saturation
occurs
at
3.4 mmol g- where the difference between the
N/Al ratios determined by TPD and by XPS is of
the order of 0.3. It appears thus that around 309/o
of the acid sites of this zeolite are weak enough to
be uncovered
at room
temperature
under
lo- Pa (operating
conditions
of the XPS
analyses).
The evolution of the concentration
of NH,
adsorbed on Lewis sites confirms and clarifies the
FTIR results. Indeed, half of the Lewis sites titrated by XPS (recall that the weakest sites are not
titrated)
are covered by the first dose of NH,
(0.35 mmol g-) which, according to calorimetry
(Fig. 7), corresponds
to adsorption
enthalpies
larger than 150 kJ mol - . Afterwards,
while
adsorption
on Briinsted sites still occurs after
the first dose, the adsorption
on Lewis :sites is
nearly
non-existent
until
the fourth
dose
(1.4 mmol g-l), and then increases again up to
saturation.
This evolution evidenced by Fig. 14
thus demonstrates the strong acidity of a sizeable
proportion
of Lewis sites, the weak acidity of the
others, and the intermediate
acidity of most
Briinsted sites.
After coverage of nearly half of the sites, the

288

A. BorPave

et al. / Microporous

Materials

II (1997)

275-291

0.8 -

0.7

1.4

2.1

2.8

3.5

4.2

4.9

5.6

6.3
mmol

N/Al (overall) XPS


N/Al (Lewis) XPS

+
--c

N/Al (Bransted) XPS


N/Al (overall)TPD

7.0
(NH&g.

Fig. 13. N/Al atomic ratios as a function of the amount of ammonia sent over LZ-Y62 measured by XPS and TPD.

B/L ratio remains constant. Thus, it appears that


there exists a sizeable proportion of Briinsted sites
among the weak sites of this LZY62 zeolite. These
sites are probably those which have three aluminum atoms as second-nearest neighbors [ 681. Their
number decreases when dealumination
increases,
which is in agreement with the evolution of the
TPD curves (Figs. 2 and 5).
The sites surrounded by fewer aluminum atoms
belong to the intermediate category of sites. As
shown by FTIR [43,67,69] (Fig. 9) they correspond to (HF) and (LF) OH groups, the last ones
being less affected by pyridine adsorption. In
agreement with the study by Chen et al. [44], the
heats of adsorption of ammonia or pyridine correponding to these medium Bronsted sites tend to
decrease with increasing steam treatment. This is
consistent with the shift of Tdesin the TPD experiments (605 K for CBV 712, 580 K for CBV 760)
and could been related to a perturbation of acidic

1:A
2-

8.25

03

0.35

0.4

0.45

0.5

N/AI

0.55

0.6

Fig. 14. Briinsted sites/Lewis sites ratios (measured


a function of the coverage (N/Al) in LZ-Y62.

0.65

0.7

by XPS) as

(HF) OH groups by the molecules a.dsorbed on


(LF) OH [43 ] or on EFAL.
Whereas most Bronsted sites corresponding to
the bridged hydroxyl groups of the structure are
of medium strength, some of them display strong
acidity, as evidenced by the FTIR and XPS data,
especially in the case of the dealuminated zeolites.
The acidity of these sites may be explained either
by an interaction of some bridged OH groups with
the Lewis sites 170,711, or by the formation of
amorphous silica-alumina
inside the pores [72],
or by an inductive effect of aluminum cations
located in the sodalite cages [73]. Although it is
difficult to establish the genuine reason for the
strong acidity of these Bronsted sites, some of our
results tend to substantiate the first hypothesis.
Indeed, the second one (formation of amorphous
silica-alumina) fails to account for the presence of
these sites in the non-dealuminated
and well crystallized LZ-Y62 zeolite. In that case the Lewis sites
must be mainly associated with framework defects
rather than EFAL. The third hypothesis, favored
by Jolly et al. [66], would imply that the aluminum
cations (strong Lewis sites) located in the sodalite
cages of the dealuminated zeolites cannot be titrated by pyridine but only by NH,. However the
proportions of Lewis sites titrated by QH,N and
by NH3 at 373 K and at 673 K follow similar
evolutions, up to experimental uncertainties.
Besides these structural Briinsted sil:es of strong
acidity which can be found in all zeolites studied
and which correspond to NH, adsorption enthalpies between 150 and 180 kJ mol - , the dealuminated zeolites present even stronger sites, cor-

A. BorPave

et al. i Microporous

responding
to the initial heats of adsorption
(between 200 and 230 kJ mol -I). Some of these
sites are probably of the Briinsted type and located
in extra-framework
aluminum-containing
debris.
These sites are associated with the IR band around
3600 cm-
[74], and are mostly found in the
spectra of the CBV 712 and CBV 720 zeolites. The
very strong Lewis sites might be associated with
extra-framework
tetrahedral
aluminum,
which
according to Peters and Wu [75] corresponds to a
chemical displacement of 53 ppm in 27A1 MAS
NMR (compared to the 60 ppm shift observed by
them for framework tetrahedral Al) and is catalytically active, contrary to the octahedral (0 ppm)
and pentacoordinated (30 ppm) aluminum. According to Batamack et al. [76], these tetrahedral
EFAL correspond to a broad line at about 40 ppm
in hydrothermaly
treated HZSM-5 (compared to
the framework Al signal at 54 ppm). The resolution
of the spectra in Fig. 1 is not good enough to
distinguish these lines. However, the broadening
of the main signal on its right-hand side in the
spectra of the dealuminated CBV samples could
be associated with these EFAL. In the opinion of
Lonyi and Lunsford [77], the strong acidity arise
from the occurrence
of cationic
species as
A1(OH)2+
or [~l$j;A;]*+ located into the b
cages. However,
as silica-alumina,
these species
are able to block other acid sites though they can
be removed by ion exchange with aqueous solutions of ammonium
hydroxide
as shown by
Stockenhuber and Lercher [ 781.
We have shown by XPS that the use of
2,6-lutidine allows a distinction between two populations of Lewis sites, only the first of which can
be titrated by this probe molecule. The first population is present in the zeolites CBV 760 (the least
crystalline) and ZF 220 (with the highest content
of extra-framework
aluminum) as well as in silica-alumina.
Notice that these two zeolites are
those whose NMR spectra present a peak at
it seems plausible
that
30 ppm. Therefore,
2,6-lutidine only titrates Bransted sites and the
Lewis sites associated with pentacoordinated
or
tetrahedral aluminum, while it cannot reach the
Lewis sites associated with octahedral and tricoordinated aluminum. This is of course only a hypothesis which needs to be checked more thoroughly.

Materials

1 I (1997)

275-291

289

5. Conclusions
The combination of bulk and surface analysis
techniques such as TPD, calorimetry,
FTIR and
XPS coupled with chemisorption
of basic probe
molecules
(ammonia,
pyridine,
lutidine)
has
allowed a more thorough physico-chemical
characterization of the acidity of a series of de.aluminated (H-USY)
and non-dealuminated
( H-Y)
faujasites. The main conclusions drawn from the
data collected in this study follow.
( 1) In terms of acid strength, these zeolites contain
three families of sites: weak, intermediate, and
strong (beyond 160 kJ mol -I).
(2) The weak sites, associated with NH, adsorption heats below 130 kJ mol -I, are mostly of
the Lewis type and close to those of amorphous silica-alumina.
They result frosm the
degradation of the crystalline structure of the
zeolites. This degradation
becomes more
important as the zeolite is more dealuminated,
since the aluminum-containing
debris become
especially concentrated at the surface of the
crystallites.
(3) Part of these weak sites are of the BrBnsted
type and correspond
to the protons surrounded by three aluminum atoms as secondnearest neighbors. These sites are numerous
in the non-dealuminated
zeolite but vanish
after thorough dealumination.
(4) The sites of intermediate strength (adsorption
heats between 130 and 150 kJ mol --) are
mostly of the BrGnsted type, and are associated with the bridged OH groups of the structure that are surrounded
by less than three
second-nearest neighbor aluminum atoms.
(5) The
strong
(heat
between
150 and
180 kJ molt ) and very strong (between 180
and 230 kJ mol-)
sites correspond
to both
structural and extra-framework
OH groups,
as well as aluminum atoms that are either
extra-framework
(cationic, tetracoordinated)
or associated with framework defects (tricoordinated). The high strength of these various
sites is probably due to interactions between
Briinsted and Lewis sites and/or neighborhood
of cationic Al species.
(6) Finally, the potential of 2,6-lutidine as a probe

290

A. BorCave

et al. / Microporous

for discriminating
between certain types of
Lewis site has been demonstrated, and ought
to be studied more thoroughly.

Acknowledgement
We are grateful
to the Conseil
dAquitaine for financial support.

RCgional

References
[I]
[2]
[3]
[4]

[5]
[6]
[7]
[8]
[9]
[lo]
[I l]
[ 121

[13]
[I41
[ 151

[16]
[17]
[18]

[ 191
[20]
[21]

J.B. Uytterhoeven,
L.G. Christner,
W.K. Hall. J. Phys.
Chem. 69 (1965) 2117.
P.A. Jacobs, J.B. Uytterhoeven,
J. Chem. Sot. Farad.
Trans. I 69 (1973) 359.
J.W. Ward, J. Catal. 9 (1967)
225; 10 (1968)
34; 11
(1968) 259.
G. Engelhardt,
D. Michel,
High Resolution
Solid State
NMR of Silicates and Zeolites. John Wiley and Sons, New
York, 1987.
A.L. Blumenfeld,
D.J. Coster, J.J. Fripiat, J. Phys. Chem.
99 (1995) 15181.
D. Coster, A.L. Blumenfeld,
J.J. Fripiat,
J. Phys. Chem.
98 (1994) 6201.
G.H. Ktihl, ACS Symp. Ser. 40 (1977) 96; J. Phys. Chem.
Solids 38 (1977) 1259.
J. Abbot, Appl. Catal. 47 (1989) 33.
V.L.
Zholobenko,
L.M.
Kustov,
B.V.
Kazansky,
E. Loeffler, U. Lohse, G. Oehman, Zeolites I 1 (1991) 132.
C. Mirodatos,
D. Barthomeuf,
J. Chem. Sot. Chem.
Commun.,
39 (1981).
Y. Yong, V. Gruver,
J.J. Fripiat, J. Catal. 150 (1994) 421.
C. Guimon,
A. Zouiten, A. Boreave, G. Pfister-Guillouzo,
P. Schulz, F. Fitoussi,
C. Quet, J. Chem. Sot. Farad.
Trans. 90 (1994) 3461.
D.A. Shirley, Phys. Rev. B 5 (1972) 4709.
J.H. Scolield,
J. Elect. Spectrosc.
Relat.
Phenom.
8
(1976) 129.
D. Freude, T. Frohlich,
H. Pfeifer, G. Scheler, Zeolites, 3
(1983) 171; D. Freude, M. Henger, H. Pfeifer, Z. Physik.
Chem. N. F. 152 (1987) 429.
V. Bosacek,
D. Freude,
T. Friihlich,
H. Pfeifer,
H. Schmiedel, J. Colloid
Interface
Sci. 85 (1982) 502.
X. Yang, J. Phys. Chem. 99 (1995) 1276.
A. Corma,
V. Fornes,
A. Martinez,
J. Sanz, Fluid
Catalytica
Cracking:
Role in Modem
Refining, Ch. 2, Am.
Chem. Sot., 1988.
D. Coster, A.L. Blumenfield,
J.J. Fripiat, J. Phys. Chem.
98 (1994) 6201.
J. Klinowski,
C.A. Fyfe. G.C. Gobbi, J. Chem. Sot. Farad.
Trans. I 81 (1985) 3003.
D. Freude,
E. Brunner,
H. Pfeifer,
D. Prager,
H.G.

Materials

I1 (1997)

275-291

Jerschkewitz,
U. Lohse, G. Dehlmann,
Che-m. Phys. Lett.
139 (1987) 325.
H. Hamdan,
Chem. Phys. Lett.
PI D. Freude, J. Klinowski,
149 (1988) 355.
1231 G. Bagnasco, J. Catal. 159 (1996) 249.
V. Forms,
F.V. Melo, J. Herrero,
Zeolites 7
]241 A. Corma,
(1987) 559.
l251 G. Zi, T. Yi, Z. Yugin, Appl. Catal. 56 (1989) 83.
Zeolites IO (1990) 307;
[261 M. Sawa, M. Niwa, Y. Murakami,
10 (1990) 532.
1271G. Seo, H.S. Jeong, S.B. Hong, Y.S. Uh, Catal. Letts. 36
( 1996) 249.
J.A. Dumesic,
Adv. Catal.
38
P81 N. Cardona-Martinez,
(1992) 149.
in Zeolites
and related
[291 H.G. Karge, L.C. Jozefowicz,
microporous
materials:
state of the art, J. Weitkamp
et al.
(Eds.), Studies in Surface Science and Catalysis,
vol. 84,
Elsevier, Amsterdam,
1994, p, 685.
[301 P.J. Andersen, H. Kung, Catalysis 11 (1995) 441.
[311 W.E. Fameth, R.J. Gorte, Chem. Rev. 95 t 1995) 615.
P.C. Gravelle,
Thermochim.
1321A. Auroux, P. Wierskowsky,
Acta 32 (1979) 165.
Y. Ben Taarit,
Thermochim.
Acta
122
[331 A. Auroux,
(1987) 63.
J. Valyn,
G.I.
Kasputin,
D.
Kalo.
[341 I. Bankos,
A.L. Klyachko,
T.R. Brueva, Zeolites 8 (1!)88) 189.
1351 A. Auroux, Y.S. Jin, J.C. Vedrine, L. Benoit, Appl. Catal.
36 (1988) 323.
K. Karim,
D. Dwyer,
Microporous
[361 M.A. Makarova,
Mater. 4 (1995) 243.
[371 C. Lee. D.J. Parillo, R.J. Gorte, W.E. Fa.rneth. J. Am.
Chem. Sot. 118 (1996) 3262.
Chakrabarty,
C.S. Handloser,
M.W.
Mosher.
[381 M.R.
J. Chem. Sot. Perkin. Trans. 2 (1972) 938.
D.L.
King,
1. Wadso,
J. Chem.
[391 C.E. Vanderzee,
Thermodyn.
4 (1972) 685.
[401 D.H. Ave. M.T. Bowers, Gas Phase Ion Chemistry, vol. 2,
Academic
Press, New York, 1979, p. 1.
J. Am. Chem.
[411 D.J. Parrillo, R.J. Gorte. W.E. Farneth,
sot. 115 (1993) 12441.
[421D.J. Parrillo, R.J. Gorte, J. Phys. Chem. 97 (1993) 8786.
T. Chevreau,
J.C. Lavalley.
Microporous
[431 S. Khabtou,
Mater. 3 (1994) 133.
J.A. Dumesic,
[441 D. Chen, S. Sharma, N. Cardona-Martinez,
V.A.
Beil, G.D.
Hodge,
R.J. Madon,
J. Catal.
I36
(1992) 392.
[451 P.E. Eberly, Jr., J. Phys. Chem. 72 (1968) 1042.
Y. Ben Taarit, Can. J. Chem. 66
[461 Z.C. Shi, A. Auroux,
(1988) 1013.
[471 Y. Mitani, K. Tsutsumi, H. Takahashi, Bull. Chem. Sot.
Jpn 56 (1983) 1917.
Ph.D.
Dissertation,
University
[481 N. Cardona-Martinez,
Wisconsin-Madison,
1989.
J.A. Dumesic,
J. Catal.
125
[491 N. Cardona-Martinez,
( 1990) 427.
1501 D. Chen, S. Sharma, I. Filimov, J.A. Dumejic, Catal. Lett.
12 (1992) 201.

A. Bortave
[51]
[52]

[53]
[54]
[55]
[56]
[57]
[58]
1591
[60]
[61]
[62]
[63]
[64]

et al. i Microporous

H. Knozinger,
P. Ratnasamy,
Catal. Rev. Sci. Eng. 17
(1978) 31.
G. Coudurier,
F. Lefebvre,
Les techniques
physiques
detude des catalyseurs,
in B. Imelik, J.C. Vedrine (Eds.),
Technip, Paris, 1988, p. 29.
J.C. Lavalley,
Catal. Today 27 (1996) 377.
L.M.
Kustov,
V.B. Kasansky,
S. Beran, L. Kubelkova,
P. Jiru, J. Phys. Chem. 91 (1987) 5247.
L. Kubelkova,
S. Beran, J.A. Lercher,
Zeolites 9 (1989)
539.
T.R. Hughes, H.M. White, J. Phys. Chem. 71 (1967) 2192.
P.A. Jacobs, B.K. Theng, J.B. Uytterhoeven,
J. Catal. 26
(1972) 191.
J.W. Ward, in J. Rabo (Ed.),
Zeolite
Chemistry
and
Catalysis, Am. Chem. Sot., Washington
DC, 1976, p. 118.
J. Datka, B. Gil, A. Kubacka,
Zeolites I5 (1995) 501.
T. Barzetti,
E. Selli, D. Moscotti,
L. Forni, J. Chem. Sot.
Faraday
Trans. 92 (1996) 1401.
A. Janin, M. Maache,
J.C. Lavalley,
J.F. Joly, F. Raatz,
N. Szydlowsky,
Zeolites I1 (1991) 391.
D. MC Queen, B.H. Chiche, F. Fajula, A. Auroux,
C.
Guimon,
F. Fitoussi, P. Schulz, J. Catal. 161 (1996) 587.
C. Defosse, P. Canesson,
J. Chem. Sot. Faraday.
Trans.
1 (1976) 2565.
R.B. Borade,
A. Adnot,
S. Kaliaguine,
J. Chem. Sot.

Materials

[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]

II (1997)

275-291

291

Faraday.
Trans.
86 (1990)
3949; J. Mol.
Catal.
61
(1990) L7.
P.V. Shertukde,
W.K. Hall, J.M. Dereppe,
G. Marcelin,
J. Catal. 139 (1993) 468.
S. Jolly, J. Saussey, J.C. Lavalley,
J. Mol. Catal. 86
(1994) 401.
M.A.
Makarova,
K. Karim,
J. Dwyer,
Microporous
Mater. 4 (1995) 243.
B. Hunger,
H. Miessner,
M.V. Szombathely,
E. Geidel,
J. Chem. Sot. Faraday
Trans. 92 (1996) 499.
J. Datka, M. Boczar, B. Gil, Colloids Surf. 105 (1995) 1.
R.A.
Beyerlien,
G.B. MC Vicker,
L.N. Yacullo,
J.J.
Ziemak,
J. Phys. Chem. 92 (1988) 1967.
A. Corma, F. Fornes, F. Rey. Appl. Catal. 59 (1990) 267.
G. Garralon,
A. Corma, V. Fornes. Zeolites 9 (1389) 84.
R. Carvajal,
P.J. Chu, J.H. Lunsford,
J. Catal.
125
(1990) 123.
U.
Lohse.
E. Loffler,
M.
Hunger,
J. Stbckner,
V. Patzelova,
Zeolites 7 (1987) 11.
A.W. Peters, C.C. Wu, Catal. Lett. 30 (1995) 171.
P. Batamack,
C. Doremieux-Morin,
J. Fraissard,
D. Freude, J. Phys. Chem. 95 (1991) 3790.
F. Lonyi, J.H. Lunsford.
J. Catal. 136 (1992) 566.
M. Stockenhuber,
J.A. Lercher,
Microporous
Mater.
3
(1995) 457.

Das könnte Ihnen auch gefallen