You are on page 1of 20

Eect of the degree and type of the dealumination method

on the structural, compositional and acidic characteristics


of H-ZSM-5 zeolites
Costas S. Triantallidis
a
, Athanasios G. Vlessidis
a
, Lori Nalbandian
b
,
Nicholaos P. Evmiridis
a,
*
a
Laboratory of Inorganic and Analytical Chemistry, Department of Chemistry, University of Ioannina, Ioannina 45 110, Greece
b
Center for Research and Technology Hellas, Chemical Process Engineering Research Institute, P.O. Box 361, 570 01 Thermi,
Thessaloniki, Greece
Received 23 October 2000; received in revised form 12 June 2001; accepted 26 June 2001
Abstract
A series of H-ZSM-5 zeolites with dierent framework Si/Al ratios were prepared by hydrothermal synthesis and
post-synthesis dealumination by various methods, i.e., HCl, steaming, steaming/HCl and ammonium hexauorosilicate
(AHFS). The degree of framework dealumination and the amount/type of the extra-framework phases formed were
greatly dependent on the method of dealumination and the severity of the experimental conditions. The hydrothermal
treatment of a parent ZSM-5 sample with Si/Al ~ 27 was very eective and produced signicant amounts of SiAl
extra-framework phases, while the reaction with HCl in relatively strongly acidic environments resulted in very low
dealumination. Treatment with AHFS was proven to be an appropriate method for preparing moderately dealuminated
H-ZSM-5 samples which are free from extra-framework Al. The relative crystallinity and the microporosity of all
dealuminated H-ZSM-5 zeolites were retained to a high degree (>80%), even for the severely steamed samples (~96%
dealumination). The morphology of the crystals/particles was not changed signicantly; however, each dealumination
method had a dierent eect on the partial breakdown of the crystals and on the formation of extra-framework/
amorphous phases with a mesoporous character. The number of acid sites that corresponded to the high-temperature
desorption peak of the ammomia-temperature programmed desorption spectra was found to be equivalent to the
framework Al (FAl) content of all H-ZSM-5 samples, irrespective of the degree of dealumination and the amount of
extra-framework phases. Theses phases had a low capability of adsorbing ammonia compared to the FAl atoms and
were the main source of acidity for the severely steamed samples which had almost no FAl. 2001 Elsevier Science
B.V. All rights reserved.
Keywords: H-ZSM-5 zeolites; Dealumination; Extra-framework phases; Crystal morphology and porosity; Acidity
1. Introduction
H-ZSM-5 zeolite is a typical example of a shape-
selective acid catalyst [1,2]. Its unique catalytic
performance is attributed to the two-dimensional
Microporous and Mesoporous Materials 47 (2001) 369388
www.elsevier.com/locate/micromeso
*
Corresponding author. Tel.: +30-651-98-404; fax: +30-651-
44-831.
E-mail address: nevmirid@cc.uoi.gr (N.P. Evmiridis).
1387-1811/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S1387- 1811( 01) 00399- 7
system of intersecting channels and to the rela-
tively low number of strong acid sites due to the
high Si/Al ratio (>10) of the framework [3]. The
chemical uniformity of the Br onsted-type acid sites
(framework hydroxyl groups) in H-ZSM-5 zeolites
with dierent framework Al (FAl) contents has
been proven in the n-hexane cracking test reaction,
where the cracking activity is linearly related to
FAl content, even down to ~20 ppm of Al [4].
Furthermore, infrared (IR) studies of highly crys-
talline H-ZSM-5 zeolites revealed the existence of
one single type of bridging hydroxyl group related
to the FAl atoms [5,6]. However, (hydro)thermal
treatment of ZSM-5 produces extra-framework
Al (EFAl) species which aect the acid catalytic
activity of dealuminated H-ZSM-5 zeolites [710].
This behavior has also been recognized and is even
more pronounced in the case of Y-type dealumi-
nated zeolites [1113]. Furthermore, it has been
shown that both the amount and the type of extra-
framework/amorphous Al or SiAl species/phases
are important in dening the acidity and catalytic
activity of the catalysts. In general, it has been
shown that mild steaming conditions result in
catalysts with enhanced cracking activity either
due to increased acid strength of the remain-
ing framework OH or due to the interaction of
the reactant alkane molecule with both the FAl
hydroxyls and the EFAl species [710,14]. Addi-
tionally, FT-IR studies of hydrothermally dealu-
minated H-ZSM-5 zeolites revealed the formation
of weakly acidic hydroxyls on EFAl species at mild
steaming conditions, which progressively disap-
pear as the severity of the treatment increases [15
17].
In larger scale reactions/processes like the
cracking of gas-oil in uid catalytic cracking
(FCC) units for the production of gasoline and
gaseous products, the performance of ZSM-5 as
additive to the main (RE)USY cracking catalyst
changes gradually as it is being dealuminated/de-
activated between the successive reaction/regenera-
tion steps in the FCC riser and regenerator,
respectively [1,18,19]. The product distribution in
the FCC process shows that the ``fresh'' H-ZSM-5
additive is capable of catalyzing the cracking of
short chain alkanes which requires high acid
strength, but after deactivation it can mainly
contribute to alkene isomerization. In a previous
work, we have shown how the decrease of the
number of acid sites for a series of dealuminated
H-ZSM-5 additives of a commercial RE-USY
catalyst aect the product distribution and prop-
erties in the cracking of gas-oil [20]. The H-ZSM-5
samples used in the referenced work were dealu-
minated by dierent methods and to dierent ex-
tents in order to closely follow the changes induced
by variation in acidity.
A number of dealumination techniques have
been developed for zeolites, especially for the Al-
rich Y-type zeolites, i.e., treatment with steam or
SiCl
4
vapor at elevated temperatures or treatment
with ammonium hexauorosilicate (AHFS), acids
(i.e., HCl), chelating agents (i.e., EDTA), etc.
[1113,2123]. Zeolite ZSM-5 is more resistant to
dealumination and relatively severe process condi-
tions are necessary to obtain highly dealuminated
samples. The most commonly used methods are
the hydrothermal dealumination (sometimes com-
bined with HCl treatment) and the treatment with
SiCl
4
vapors at elevated temperatures [9,1416,24
30], while the use of mineral acids alone has also
been studied [31]. In order to simulate the dealu-
minated/deactivated state of the ZSM-5 additive
in the FCC catalyst, we applied in this work
relatively severe steaming conditions for the de-
alumination of the ZSM-5 samples, while mildly
steamed samples were also prepared to follow the
progressive changes in the acidic properties. Treat-
ment with HCl was also applied either for direct
dealumination or after steaming in order to re-
move the EFAl phases formed. In addition to
these methods, the dealumination of zeolite ZSM-5
with AHFS has also been tested. A limited num-
ber of studies have focused on the dealumination
of ZSM-5 with AHFS [20,32,33], since it is a rela-
tively mild dealumination agent. Its main advan-
tage, however, is that it produces samples almost
free of EFAl species, and this could be benecial
in avoiding the blockage of the acid sites from
EFAl species which can be located either on the
outer surface of the crystals or inside the narrow
channels of ZSM-5 zeolites.
The present study aims at a better understand-
ing of the unique eects that each dealumination
method induces on the compositional, structural
370 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
and acidic characteristics of dealuminated H-
ZSM-5 zeolites and rationalize their catalytic be-
havior in the FCC process as well as in other, acid
catalyzed reactions. The composition of the crys-
talline and extra-framework/amorphous phases of
the samples has been determined by the combina-
tion of chemical analysis and
29
Si
27
Al MAS-NMR
data, while structural and porosity characteristics
were studied by XRD, FT-IR, nitrogen adsorp-
tion/desorption, SEM and TEM experiments. The
acidic characteristics of the H-ZSM-5 samples were
tested by ammonia-temperature programmed de-
sorption (TPD) measurements and correlated
with the composition of the crystalline and amor-
phous phases present in the dealuminated sam-
ples.
2. Experimental section
2.1. Preparation of the samples
The ZSM-5 zeolite samples with dierent Si/Al
ratios were synthesized in this laboratory, based
on the method introduced by Argauer and Landolt
[34], and their crystallographic purity was deter-
mined by X-ray diraction. The as-synthesized
samples were calcined under a ow of dry air at
600C for 3 h in order to combust the organic
template. They were subsequently treated with a
0.25 N HCl acid solution at 80C for 24 h to ion-
exchange Na

by H

and produce the H-form


of the zeolites (samples ZM-1 to ZM-4, Table 1)
or they were ion-exchanged twice with a 1 M
NH
4
NO
3
solution at 80C for 24 h to produce
NH
4
ZSM-5 zeolites.
A calcined ZSM-5 sample (Si/Al = 27) was
treated with an excess of a 1.5 M HCl acid solution
at 90C for 24 h in order to induce framework
dealumination (sample DZM-5). The NH
4
ZSM-5
sample with an Si/Al ratio 27 was treated with
AHFS at 80C for 6 h. Two samples were prepared
by this method; the rst one was treated twice with
a 0.1 N AHFS solution using an amount of AHFS
that matches the FAl content of the zeolitic frame-
work (sample DZM-6); the second sample was
treated ve times with an excess of a 0.5 N AHFS
solution (sample DZM-7). These samples were
transformed to the H-forms by calcination at
530C for 2 h under a ow of dry nitrogen.
The NH
4
ZSM-5 samples were hydrothermally
treated at dierent temperatures (570C, 660C,
730C, 790C), for 6 h under a ow of a nitrogen
steam mixture at a ow rate of 30 ml min
1
and a
steam partial pressure of 90.4 kPa, in order to vary
the severity of steaming and the degree of dealu-
mination (samples DZM-8, DZM-9, DZM-10,
DZM-11). The steamed sample DZM-10 was fur-
ther treated with a 0.2 N HCl solution at 80C for
24 h in order to extract the EFAl species formed
(sample DZM-12).
All the parent and dealuminated samples were
dried at 120C in air for 6 h and were nally stored
over saturated MgCl
2
solution to equilibrate with
water vapor. The parent and dealuminated sam-
ples that were not hydrothermally treated are
usually designated in the text as ``fresh'' samples in
contrast to the steamed samples. All samples pre-
pared and the treatment methods applied are listed
in Table 1.
2.2. Composition and structure characterization
The chemical analysis procedure of the samples
involved a digestion step with 10% HCl, to extract
the Al
3
and Na

into the acid aqueous solution;


silica was then determined gravimetrically from
the solid residue; aluminum was determined in
solution titrimetrically by EDTA complexation;
the sodium content was measured by ame pho-
tometry, and the water content was estimated from
the weight loss after calcining the samples at
550C for 24 h.
The X-ray powder diraction patterns were
obtained on a Siemens D-500 automated dirac-
tometer (CuKa radiation, k = 1:5418

A), in the 2h
range of 535 and at a scanning rate of 2 min
1
.
The zeolites were checked for crystallinity by
comparing the intensity of the (0 5 1) peak of the
dealuminated samples with that of the parent
ZSM-5 zeolite which was considered to be 100%
crystalline. Mid-IR spectra were recorded at 2
cm
1
resolution on a Perkin Elmer 1650 FT-IR
spectrometer using the KBr-pellet technique (1
wt.% zeolite in a KBr matrix). High-resolution
solid state
29
Si MAS NMR spectra were recorded
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 371
using a Bruker MSL 400 spectrometer working at
79.49 MHz, with a pulse width of 12 ls, a pulse
delay of 100 s, a spinning rate of 4 kHz and 1000
scans. The
27
Al MAS NMR experiments were run
on a Varian Innity plus AS400 spectrometer
at 104.26 MHz, with a pulse width of 0.5 ls, a
radiofrequency eld strength of 50 G, a pulse delay
of 0.5 s, a spinning rate of 4 kHz and 85,000 scans.
29
Si shifts were referenced to TMS by using an
external sample of sodium 2,2-dimethyl-2-silapen-
tane-5-sulfate (SDS) which gave a single peak at
0 ppm as in the case of TMS, while a 1 M aqueous
Al(NO
3
)
3
solution was used as standard reference
for
27
Al. All samples were dried at 120C in air for
6 h and were nally stored over saturated MgCl
2
solution to equilibrate with water vapor before
testing.
Specic surface area (SSA) and porosity char-
acteristics of the samples were determined from
adsorption/desorption isotherms of nitrogen which
were obtained at 77K on an Automatic Volumet-
ric Sorption Analyzer (Autosorb-1, Quantachrome).
Prior to the determination of the adsorption iso-
therms the samples were evacuated overnight at
500C under 1:0 10
6
mbar vacuum.
Transmission electron microscopy (TEM) im-
ages were obtained on a JEOL JEM-100CX II
microscope with a CeB
6
lament and an acceler-
ating voltage of 120 kV. The TEM samples were
prepared by evaporating one drop of a powdered
zeoliteEtOH suspension (after sonication) onto
a carbon-coated, holey lm supported on a 3
mm, 300 mesh copper grid. The morphology of
the zeolitic particles was also examined by scan-
Table 1
Compositional and structural characteristics of the H-ZSM-5 samples (chemical analysis, XRD, FT-IR)
Sample Treatment Chemical analysis Si/Al XRD FT-IR Frequency
of bands
a
Composition (wt.%) Si and Al con-
tent (mg-
atoms g
1
)
(0 5 1)
peak
2h (de-
grees)
Rela-
tive
cryst.
(%)
b
SiO
2
(1)
Al
2
O
3
(0.2)
Na
2
O
(0.03)
H
2
O
(0.3)
Q
TSi
c
(0.01)
Q
TAl
(0.04)
m
1
(cm
1
)
(1)
m
2
(cm
1
)
(1)
ZM-1 Parent 84.77 5.75 0.20 9.16 14.11 1.13 12.5 23.11 ~97.5 1091.3 793.8
ZM-2 Parent 90.34 2.82 0.08 6.54 15.04 0.55 27.2 23.12 100 1099.9 797.2
ZM-3 Parent 92.22 2.13 0.15 5.30 15.35 0.42 36.7 23.16 100 1094.5 794.7
ZM-4 Parent 98.91 0.08 0.06 0.77 16.46 0.02 >10-
00
23.10 100 1104 799.4
DZM-5 HCl 90.41 2.62 0.13 6.68 15.05 0.51 29.3 23.13 94
d
1100.8 797.7
DZM-6 (NH
4
)
2
SiF
6
93.33 1.95 0.14 4.44 15.53 0.38 40.6 23.08 93 1103.6 799.1
DZM-7 (NH
4
)
2
SiF
6
95.24 1.63 0.10 2.94 15.86 0.32 49.6 23.10 92 1104.3 799.7
DZM-8 Steamed,
570C
94.40 2.95 0.08 2.46 15.71 0.58 27.2 23.08 89 1106 800.1
DZM-9 Steamed,
660C
94.98 3.01 0.07 1.82 15.81 0.59 26.8 23.08 84 1106.9 801.4
DZM-10 Steamed,
730C
96.00 3.05 0.07 0.73 15.98 0.60 26.7 23.08 80 1107.7 802
DZM-11 Steamed,
790C
96.35 3.08 0.07 0.42 16.04 0.60 26.5 23.10 82 1107.9 802.6
DZM-12 DZM-10/HCl 96.11 2.33 0.04 1.34 16.00 0.46 35.0 23.11 84 1107.3 802.5
a
Frequencies of the main asymmetric stretch (m
1
) and the main symmetric stretch (m
2
) vibrations.
b
Relative crystallinity of the samples using the (0 5 1) peak intensity from XRD patterns and considering the parent zeolite samples
100% crystalline except for sample ZM-1 which possesses some amorphous material as explained in the text.
c
Q
TSi
and Q
TAl
= total Si and Al contents per gram hydrated sample, where sample = crystalline zeolite plus amorphous material.
d
The parent zeolite for all dealuminated samples (DZM-5 to DZM-12) was sample ZM-2.
372 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
ning electron microscopy (SEM) images which
were taken on a JEOL JSM-6300 scanning mi-
croscope.
2.3. Acidity characterization
The number, strength and distribution of acid
sites were tested by TPD of ammonia experiments
which were performed on a conventional appara-
tus consisting of a cylindrical quartz micro-reac-
tor, a vertical furnace and a thermal conductivity
detector (TCD) of a Shimadzu GC-8A gas chro-
matograph. Typically, a 0.2 g zeolite sample (3 mm
thick bed) was placed in the reactor and was
heated up to 530C with a temperature program of
5Cmin
1
under He ow (5060 ml min
1
); at this
temperature the sample was further heated for 3 h.
Adsorption of dry ammonia (Merck, water free)
took place at 100C, in a static system for 1 h at
1.5 bar ammonia pressure. Prior to desorption the
samples were ushed with He for 1.5 h at 100C
to remove the physically adsorbed ammonia.
Desorption of ammonia was done at a rate of
10Cmin
1
from 100C up to 700C under a He
ow (50 ml min
1
). The desorbed ammonia was
detected on the TCD and then trapped in a 0.01 N
HCl aqueous standard solution. The trapped am-
monia was determined by titrating the excess of
HCl in a 0.01 N standard NaOH solution.
3. Results and discussion
3.1. Total Si and Al contents chemical analysis
The chemical composition of the parent and
dealuminated H-ZSM-5 samples is given in Table
1. The chemical analysis of the samples can pro-
vide the total Si and Al contents, including both
framework and extra-framework species, which
was used to calculated the bulk (total) Si/Al ratio.
The total Si/Al ratio for the parent zeolite samples
(ZM-1 up to ZM-4) ranges from 12.5 to >1000
(ZM-4 was synthesized without any Al source)
while for the samples dealuminated by HCl and
AHFS it ranges from 29 to 50. On the other hand,
the total Si/Al ratio for all the steamed samples is
almost constant and similar to that of the parent
zeolite (~27) since the extra-framework species
formed during the hydrothermal-treatment are not
leached out of the solid, except for the steamed/
HCl-treated sample (DZM-12) which has a bulk
Si/Al ratio of 35.
3.2. Structural properties and relative crystallinity
XRD and FT-IR spectroscopy
The structural and crystallinity characteristics
of the samples were studied by XRD (Fig. 1a) and
FT-IR spectroscopy (Fig. 1b). All the parent
samples are highly crystalline as it is revealed from
their XRD patterns, and they are almost free
from extra-framework/amorphous phases, except
Fig. 1. (a) XRD patterns and (b) IR spectra of the parent
sample ZM-2 and the steamed sample DZM-11.
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 373
for ZM-1; this sample possesses some amorphous
material as indicated by the octahedral Al ob-
served in the
27
Al MAS NMR spectra and from
the presence of irregularly shaped particles of ex-
tra-framework phases found in the SEM images,
as discussed below. All the dealuminated samples
(DZM-5 up to DZM-12) were prepared from the
parent sample ZM-2, and their relative crystallin-
ity (compared to ZM-2) was estimated on the ba-
sis of the intensity of the (0 5 1) peak in the XRD
pattern which is characteristic of the ZSM-5
structure. Relative crystallinity data and the posi-
tion of the (0 5 1) peak (at ~23 2h) for all the
samples are given in Table 1. It can be seen that all
the dealuminated samples retain a relative crys-
tallinity equal to or higher than 80%; furthermore,
the position of the (0 5 1) peak is slightly aected
by dealumination (framework dealumination in-
creases from DZM-5 to DZM-12 as shown below).
Consequently, the dimensions (a; b; c) and the
volume of the unit cell (not shown here) do not
change signicantly and systematically with de-
alumination of the parent ZSM-5 sample; this can
be attributed to the relatively small changes that
occur in the unit cell of a ZSM-5 zeolite, even
when all the Al is removed from its framework,
since the parent sample usually does not have
more than 34 Al atoms per unit cell (and 9293 Si
atoms). Similar results for the dealumination of
large ZSM-5 crystals have been previously re-
ported [27].
In contrast to the unit cell parameters, the IR
frequencies of the main asymmetric stretch vi-
bration (at ~10911108 cm
1
) and of the main
symmetric stretch vibration (at ~790802 cm
1
)
are systematically aected by dealumination.
Straightforward relationships between the fre-
quencies of the vibrations and the FAl content of
all the dealuminated samples were derived (as
shown below), irrespective of the method of de-
alumination and the extra-framework phases pre-
sent in the samples.
3.3. Framework Al content solid state MAS NMR
spectroscopy
29
Si and
27
Al solid state MAS NMR spectro-
scopy is usually applied for the determination of
the FAl and EFAl content of zeolites, as well as for
examining the coordination state of EFAl species
[10,15,16,23,24,3544]. In the present study, the
FAl content of the dealuminated ZSM-5 samples
was determined based on the integrated intensity
(area) of the peak at ~54 ppm (assigned to tetra-
hedrally coordinated Al) in the
27
Al MAS NMR
spectra by using an appropriate calibration line as
explained below.
27
Al MAS NMR spectra for the parent material
ZM-2 and samples dealuminated in dierent
manners are shown in Fig. 2. The parent ZM-2
sample (as well as the rest of parent samples)
shows a strong peak at ~54 ppm due to tetra-
hedrally coordinated FAl and a very weak peak at
~0 ppm due to octahedrally coordinated EFAl;
these few EFAl species in the parent samples were
formed during the calcination of the as-synthe-
sized zeolites for the combustion of the organic
template and the subsequent treatment with dilute
HCl solution to produce the H

-forms of the
samples. The eect of these treatments is stronger
on the Al-rich ZM-1 parent sample (stronger peak
due to EFAl species), although some amorphous
unreacted material was also observed in the as-
synthesized ZM-1 sample, indicating the diculty
of synthesizing Al-rich (Si/Al ~10) ZSM-5 zeolites.
The spectra of the AHFS-dealuminated sample
DZM-7 shows only the peak at ~54 ppm, sup-
porting the idea that AHFS-dealumination pro-
duces EFAl-free samples, as in the case of Y-type
zeolites [11,23,45].
The spectra of all the steamed samples show the
peak at ~54 ppm and a broad peak at ~0 ppm due
to octahedrally coordinated EFAl. The intensity of
the peak at ~54 ppm decreases progressively with
increasing severity of the hydrothermal treatment
(higher steaming temperature), while the peak at
~0 ppm decreases slightly instead of increasing
as a consequence of the formation of more EFAl
species. Furthermore, treatment of the steamed
sample DZM-10 with dilute HCl solution for the
extraction of the EFAl species leads to an increase
of the peak at ~0 ppm (although some EFAl
species were removed from the sample as shown
below), while the 54 ppm peak was not aected
signicantly. These results are in accord with pre-
viously reported data on the presence of ``NMR-
374 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
invisible'' EFAl in steamed ZSM-5 and Y zeolite
samples [10,15,16,38,43,44]. The chemical nature
of these EFAl species was assigned to oligomeric
alumina with a low degree of hydration/hydroxy-
lation which is dicult to be ion-exchanged, i.e. by
NH

4
ions. However, rehydration or complexa-
tion of these species with acetylacetone can make
them ``NMR-visible'' in the form of Al(H
2
O)
3
6
or
Al(acac)
3
complexes [10,43,44]. The broad peak/
shoulder at ~3040 ppm which was previously
observed in mildly steamed ZSM-5 [10,16] and
assigned to either tetrahedrally coordinated EFAl
species or 5-coordinated EFAl species, in analogy
to steamed Y-zeolite samples [11,22,38,4245], was
not clearly identied in the steamed samples of this
work; the very weak peak/shoulder observed at
~45 ppm could hardly be assigned to EFAl spe-
cies, and furthermore its intensity was very low
compared to the peaks of the tetrahedral FAl
atoms and the octahedral EFAl species. The rela-
tively harsh steaming conditions applied in this
work and the low FAl content of the parent
sample (Si/Al ~ 27) may account for the absence
of the peak at 3040 ppm which is related to Al
species in an intermediate, extra-framework envi-
ronment between the framework tetrahedral Al
and the extra-framework octahedral Al.
The intensities (areas) of the peak at 54 ppm
due to framework tetrahedral Al, estimated after
deconvolution of the experimentally determined
27
Al MAS NMR spectra of the samples using
mixtures of Gaussian and Lorentzian curves, are
given in Table 2. The calibration line for the
determination of the FAl content of the dealu-
minated samples was estimated based on the par-
ent sample ZM-1 and the dealuminated samples
DZM-6 and DZM-7. The FAl content of sample
ZM-1 was determined based on its
29
Si MAS
NMR spectra (shown in Fig. 3); ZM-1 is a ZSM-5
sample relatively rich in Al, and this enabled us to
estimate the framework Si/Al ratio by using the
equation Si=Al = I
(total)
=0:25 I
Si(1Al)
[16,35,37,39].
From the relative intensities of the deconvoluted
Gaussian peaks of the spectra in Fig. 3 and by
using the above equation, the framework Si/Al
ratio of ZM-1 was estimated to be 22.6. The
Fig. 2.
27
Al MAS NMR spectra of the parent sample ZM-2, the AHSF-dealuminated sample DZM-7 and the steamed samples
DZM-9 to DZM-12.
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 375
framework Si (FSi) and FAl contents of ZM-1
were then calculated based on the following
equations:
Q
FAl
= (N
FAl=u:c:
=MW) Y
Q
FSi
= (N
FSi=u:c:
=MW) Y
(1)
where Q
FAl
and Q
FSi
are the FAl and FSi contents,
respectively (in mg-atoms per gram hydrated solid
sample, where ``sample'' denotes the crystalline
zeolite plus amorphous material); N
FAl
/u.c. and
N
FSi
/u.c. are the number of FAl and FSi atoms per
dehydrated unit cell, respectively; N
FAl=u:c:
= 4:06,
N
FSi=u:c:
= 96 N
FAl=u:c:
= 91:94 (framework Si/
Al = 91:94/4:06 = 22:6, from
29
Si MAS NMR);
u.c.: unit cell; MW: weight of the dehydrated unit
cell; Y = (100 wt:% H
2
O of sample)=100 (Ta-
ble 1).
The FAl content of the AHFS-dealuminated
samples DZM-6 and DZM-7 can be considered to
be identical with the total Al (TAl) content (de-
termined by chemical analysis, Table 1) since the
27
Al MAS NMR spectra (Fig. 2) revealed the
absence of any EFAl in these samples. The FAl
contents determined (per g sample) for ZM-1,
DZM-6 and DZM-7 were then correlated to the
area of the peak at ~54 ppm of the
27
Al MAS
NMR spectra (Table 2), and a straight line with a
very good correlation coecient (r
2
= 0:9998) was
derived passing through the origin. From the in-
tercept and the slope of the above calibration line
the following equation was derived which was used
for the estimation of the FAl content per gram of
sample, from the area of the 54 ppm peak of the
27
Al MAS NMR spectra of the rest of the parent
and dealuminated samples:
Q
FAl
= (I 0:119)=46:94 (2)
where Q
FAl
is the FAl content (in mg-atoms per
gram hydrated solid sample), I is the intensity of
the 54 ppm peak of the
27
Al MAS NMR spectra
Fig. 3.
29
Si MAS NMR spectra of the parent sample ZM-1
(experimental spectra and deconvoluted peaks).
Table 2
FAl content of the H-ZSM-5 samples
Sample Framework Al content Degree of de-
alumination
a
(%)
Intensity of
27
Al-NMR peak at 54
ppm (arb. units 10
7
)
Q
FAl
(mg-atoms g
1
)
sample (0.04)
Q
FAl
b
(mg-atoms g
1
)
crystalline zeolite (0.04)
ZM-1 30.0 0.64 ~0.66
ZM-2 25.0 0.54 ~0.54
ZM-3 18.6 0.40 ~0.40
ZM-4 0.8 0.02 ~0.02
DZM-5 22.8 0.49 0.52 4
DZM-6 17.7 0.38 0.41 24
DZM-7 14.7 0.32 0.35 35
DZM-8 9.0 0.19 0.21 61
DZM-9 4.2 0.09 0.11 80
DZM-10 1.6 0.04 0.05 91
DZM-11 0.6 0.02 0.02 96
DZM-12 1.6 0.04 0.05 91
a
Based on the values of Q
FAl
per gram crystalline zeolite with reference to the parent sample ZM-2.
b
Q
FAl
per gram crystalline zeolite = (Q
FAl
per gram sample) (1=C), where C = (% relative crystallinity)=100 (Table 1).
376 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
(peak area in arb. units) and 0.119 and 46.94 are
the intercept and slope, respectively, of the cali-
bration line. The estimated FAl content per gram
sample, the estimated FAl content per gram of
crystalline zeolite and the degree of framework
dealumination for the dealuminated samples, are
given in Table 2. By considering a 10% error in the
determination of the intensity of the 54 ppm peak
of sample DZM-6, the error in the estimation of
the FAl content by Eq. (2) is 0.04 mg-atoms
per gram hydrated sample, which is similar to the
estimated error for the determination of the Al
content by chemical analysis (Table 1).
From the data in Table 2 it can be seen that
dealumination by HCl acid solution under the
relatively strongly acidic conditions applied for
sample DZM-5 is not an eective method for
ZSM-5 zeolites (4% framework dealumination
compared to the parent sample ZM-2), in accord
with previously reported data [31]. On the other
hand, the hydrothermal treatment is very eective;
dealumination increases rapidly with the steam-
ing temperature (when keeping all other condi-
tions constant) and reaches almost 96% at 790C
(DZM-11). In addition, mild treatment of the
steamed sample DZM-10 with HCl acid solution
does not aect the FAl content and removes a
small amount of EFAl species, as can be seen from
the data in Tables 1 and 3.
In the case of AHFS, a two-fold treatment of
ZM-2 with a total amount of AHFS which cor-
responds to 100% of its framework dealumination,
resulted in only 24% dealumination (DZM-6).
Extensive treatment with an excess of AHFS re-
sulted in 35% dealumination (DZM-7). However,
the advantage of this method is that no EFAl
phases are formed, as discussed below. In a recent
work, Han et al. [32], showed that dealumination
of ZSM-5 extrudates by AHFS under static di-
gestion conditions leads to selective removal of the
surface acidity, while the total number of acid sites
measured by NH
3
-TPD and hexane cracking was
reduced by approximately 30%. The authors sug-
gested that this limited dealumination was due to
the static system applied which could not favor the
diusion of the SiF
2
6
anion inside the ZSM-5
channels. The results of the present work, based on
27
Al MAS NMR and chemical analysis data, show
that the degree of ZSM-5 framework dealumi-
nation by AHFS does not increase signicantly
under stirring conditions, provided that the frame-
work Si/Al of the parent ZSM-5 samples in both
works was ~30. In a previous work on the de-
alumination of Y-zeolite by AHFS [23], we showed
that AHFS selectively removes the FAl atoms that
contribute to acid sites of weak or medium acid
strength, leaving the strong acid sites almost unaf-
fected up to ~50% dealumination of the Y-zeolite
framework. Furthermore, the dealumination ac-
tivity of AHFS was signicantly decreased after
50% dealumination, resulting in Y-zeolites with
very low crystallinity. Based on the results on both
Y and ZSM-5 type zeolites, it can be suggested
that the reaction of the SiF
2
6
anions and the FAl
atoms is mainly dependent on the number (den-
sity) of the FAl atoms and how this aects the
stability of the zeolite framework.
3.4. Formation of extra-framework phases in the
dealuminated H-ZSM-5 samples
The dierence between the values of Q
TAl
(mg-
atoms of total Al per gram sample, Table 1) and
Q
FAl
(mg-atoms of FAl per gram sample, Table 2)
gives the (EFAl) content of the samples. These
data are given in Table 3, together with compo-
sitional data of the extra-framework/amorphous
phases formed in each sample, i.e., the contents of
extra-framework Si and Al expressed as %SiO
2
and %Al
2
O
3
and the Si/Al ratio of the extra-
framework/amorphous phases. In addition, the
27
Al MAS NMR spectra provided with infor-
mation on the coordination of the Al atoms in
framework and extra-framework environments, as
discussed above. From the data in Table 3 it can
be seen that almost half of the TAl content of the
parent sample ZM-1 is on extra-framework sites.
The total Si content of this sample (determined
by chemical analysis) was found similar to the
framework Si content (determined by
29
Si MAS
NMR); thus the small amount of extra-frame-
work/amorphous material that exists in ZM-1
consists mainly of Al (~2.5% Al
2
O
3
). The EFAl
content of the other parent samples ZM-2, ZM-3
and ZM-4 and of the HCl-dealuminated sam-
ple DZM-5 is even lower than that of ZM-1, viz.
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 377
below the estimated experimental error. The
presence of EFAl species in the parent samples was
conrmed by the small peak at ~0 ppm in the
27
Al
MAS NMR spectra which corresponds to octa-
hedrally coordinated EFAl; the origin of these
EFAl species was discussed above.
By comparing the Q
TAl
and Q
FAl
values of
the AHFS-dealuminated samples DZM-6 and
DZM-7 it can be seen that they are the same and
practically no EFAl species were formed during
this treatment, as also conrmed by the
27
Al
MAS NMR spectra. However, a small amount
of amorphous material was present in these sam-
ples which mainly consisted of Si species. On the
other hand, the steamed samples possess signi-
cant amounts of EFAl species relative to their TAl
content; the amount of these species was found
to increase with increasing steaming temperature
(Table 3), although this was not shown in the
27
Al
MAS NMR spectra because of the ``NMR-invisi-
ble'' EFAl species, as discussed above. The ma-
jority of the EFAl species of the steamed samples
was found in octahedral coordination; however,
by applying relatively mild steaming conditions
(~350540C) the progressive transformation of
the EFAl from an extra-framework tetrahedral
environment to the octahedral coordination can
be monitored by
27
Al MAS NMR spectroscopy
[10,16,26]. Treatment of the steamed sample
DZM-10 with dilute HCl solution removed only
25% of the EFAl species, indicating that the
greater part exists as Al or SiAl oligomers which
are not easily ion-exchangeable, in accord with the
27
Al MAS NMR data regarding the presence of
``NMR-invisible'' EFAl species in the steamed
samples. The Si/Al ratio of the amorphous phases
of all the steamed samples is ~45 and 5.5 for the
HCl/-extracted steamed sample, indicating that
there is almost an equilibrium in the removal of Si
and Al atoms from the crystalline framework at
the dierent steaming conditions applied for these
samples.
3.5. Correlation of the framework Al content with
the infrared stretching vibrations
In addition to MAS NMR spectroscopy,
straightforward relationships between the unit cell
size (XRD) and the frequencies of certain stretch-
ing vibrations (IR) of Y-type zeolites have been
previously reported and can be used for the de-
termination of the FAl content of samples with
Table 3
Composition of extra-framework phases
Sample EFAl content Q
EFAl
a
(mg-atoms g
1
)
(0.04)
Si and Al content of extra-framework phases (%)
Amorphous material
b
Al
2
O
3
c
(0.2) SiO
2
d
(1) Si/Al
ZM-1 0.49 2.5 2.50
ZM-2 ~0(0.01) ~0(0.05)
ZM-3 ~0(0.02) ~0(0.10)
ZM-4
DZM-5 ~0(0.02) 6 ~0(0.10) 5.9
DZM-6 7 7.0
DZM-7 8 8.0
DZM-8 0.39 11 1.99 9.0 3.8
DZM-9 0.50 16 2.55 13.5 4.5
DZM-10 0.56 20 2.85 17.2 5.1
DZM-11 0.58 18 2.96 15.0 4.3
DZM-12 0.42 16 2.14 13.9 5.5
a
Estimated from the dierence Q
TAl
Q
FAl
(Tables 1 and 2, respectively).
b
Estimated from % relative crystallinity (XRD) of the dealuminated samples DZM-5 to DZM-12 (Table 1); for the parent sample
ZM-1 the amorphous phase consists mainly of Al which has been estimated as %Al
2
O
3
.
c
The extra-framework %Al
2
O
3
of the samples has been estimated from the values of Q
EFAl
.
d
The extra-framework %SiO
2
of the samples has been estimated from the dierence (% amorphous material extra-framework
%Al
2
O
3
).
378 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
framework and extra-framework phases [4648].
In order to test the application of similar rela-
tionships for the dealuminated ZSM-5 samples,
the determined FAl contents of the parent (ZM-2)
and all the dealuminated samples were correlated
to the main asymmetric stretching vibration at
~10911108 cm
1
(internal vibrations of the TO
4
tetrahedra-structure insensitive) and to the main
symmetric stretching vibration at ~790802 cm
1
(external linkages between the tetrahedra-structure
sensitive). The resulting curves were straight lines
with good correlation coecients (0.975), irre-
spective of the method of dealumination and the
amount/type of EFAl species present in the sam-
ples, as can be seen in Fig. 4a and b. From the
intercept and the slope of the straight lines in Fig.
4a and b the following equations were derived
which can be used for the estimation of the FAl
content per gram of samples:
Q
FAl
= (1108:3 m
1
)=14:48
Q
FAl
= (802:5 m
2
)=9:76
(3)
where Q
FAl
is the FAl content (in mg-atoms per
gram hydrated solid sample) and m
1
and m
2
are
the frequencies of the main asymmetric and main
symmetric stretching vibrations of IR spectra
(Table 1). The estimated error in the determination
of the IR frequencies was found to be 1 cm
1
,
which corresponds to ~0.06 (for m
1
) and 0.10
(for m
2
) mg-atoms FAl per gram hydrated sample.
3.6. Adsorption and porosity characteristics N
2
adsorption/desorption
Adsorption and porosity characteristics of the
parent and dealuminated samples derived from
nitrogen adsorption/desorption isotherms at 77 K
are given in Table 4. The relative isotherms for
the parent ZM-2 and ZM-4 samples, the AHFS-
dealuminated sample DZM-7 and the steamed sam-
ple DZM-11 are shown in Fig. 5. The adsorption
isotherm of ZM-2 is typical for an Al-rich crystal-
line/microporous H-ZSM-5 sample with relatively
small spheroid particles (~0.52 lm) which consist
of smaller individual crystallites, as it was revealed
by the SEM and TEM images (shown below). On
the other hand, ZM-4 is an Al-free highly crystal-
line sample with relatively large single crystals (~5
10 lm) of rectangular parallelepiped shape. The
type of isotherm that it exhibits is also well known
and has been a matter of detailed study in the past
leading to the conclusion that the two plateaus of
the isotherm (below 0.1 and above 0.2p=p
0
) cor-
respond to two dierent states of the adsorbed
nitrogen, a liquid-like and a solid-like phase, re-
spectively [4952]; the explanation of close packing
of the nitrogen molecules at the intersections of the
channels has also been suggested [53]. The shape
and position of the observed hysteresis loop were
found in previous work to be dependent on the
size of the ZSM-5 crystals and the content of FAl;
Fig. 4. Correlation of the frequencies of the main asymmetric
stretching vibration (a) and the main symmetric stretching vi-
bration (b) of the IR spectra with the FAl-content for the
parent sample ZM-2 and all the dealuminated samples.
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 379
larger crystals and a lower Al content resulted in a
steep increase of the adsorption isotherm and a well-
dened hysteresis loop at relative pressures lower
than 0.2p=p
0
.
Table 4
Surface area and porosity characteristics of the H-ZSM-5 samples
Samples SSA
a
(m
2
g
1
) Micro-area
b
(m
2
g
1
)
External area
b
(m
2
g
1
)
Micro-pore
volume
b
(cm
3
g
1
)
Total pore
volume
c
(cm
3
g
1
)
Meso- macro-
pore volume
(cm
3
g
1
)
ZM-1 387 347 40 0.142 0.194 0.052
ZM-2 416 331 85 0.136 0.231 0.095
ZM-3 429 331 98 0.141 0.232 0.091
ZM-4 386 349 37 0.167 0.209 0.042
DZM-5 404 319 85 0.131 0.227 0.096
DZM-6 397 323 74 0.140 0.228 0.088
DZM-7 378 303 75 0.142 0.255 0.113
DZM-8 371 315 56 0.148 0.223 0.075
DZM-9 367 308 59 0.148 0.227 0.079
DZM-10 374 285 89 0.135 0.240 0.105
DZM-11 364 281 83 0.129 0.222 0.093
DZM-12 385 303 82 0.141 0.238 0.097
a
Multipoint BET surface area.
b
t-plot method.
c
At P=P
0
~ 0:99.
Fig. 5. Nitrogen adsorption/desorption curves for the parent samples ZM-2 and ZM-4, the AHSF-dealuminated sample DZM-7 and
the steamed sample DZM-11.
380 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
Although the eect of the FAl content on the
shape of the nitrogen adsorpition/desorption iso-
therms can be recognized in both the dealuminated
samples DZM-7 and DZM-11 (Fig. 5), it is clear
that the isotherms of DZM-11 are dierent from
those of DZM-7 and ZM-4. In the mildly AHFS-
dealuminated sample DZM-7 the observed step in
the adsorption branch and the relative hysteresis
loop have developed to a small degree and are
located at a relative pressure below 0.2p=p
0
; how-
ever, a small increase of the total pore volume due
to a higher macropore volume can be observed in
this sample compared to the parent sample ZM-2
(Table 4). On the other hand, the severely dealu-
minated, steamed sample DZM-11, which has prac-
tically the same Al content as the sample ZM-4
(almost Al-free), exhibits a broader step in the
adsorption branch and a more pronounced hys-
teresis loop between 0.1 and 0.4p=p
0
. However, the
total pore volume of this sample, like the one of
the other steamed samples, is similar to that of the
parent zeolite (Table 4), indicating that this hys-
teresis loop could not be assigned to any sig-
nicant, ordered mesoporosity that might have
developed after steam dealumination of the parent
ZSM-5 sample (the absence of a Bragg peak due to
ordered mesoporosity in the low-angle region of
the XRD patterns of the steamed samples support
the above nitrogen adsorption results). In general,
from the data of Table 4 it can be seen that the
SSA of the parent sample ZM-2 does not decrease
by more than ~13% in the most strongly dealu-
minated samples (DZM-11), while the highest in-
crease in the total pore volume, which has been
observed in the AHFS-treated sample DZM-7,
does not exceed 20% of the total pore volume of
the parent sample ZM-2. These data are in accord
with the high relative crystallinity (XRD) of the
dealuminated samples and are further supported
by the TEM images of the crystals/particles, as
reported below.
3.7. Crystal/particle morphology and structural
characteristics transmission electron microscopy
The crystal/particle morphology and structural
characteristics regarding the pores and cages of
zeolites or other micro- and mesoporous materi-
als can be examined by the use of TEM [41,54,
55]. Furthermore, the changes induced to the
crystalline, microporous structure of zeolites after
post-synthesis dealumination by steaming or other
methods, can be visualized more easily by the use
of TEM; related data has been reported for de-
aluminated Y-type zeolites [56,57], ferrierites [58],
mordenites [59], and mazzites [60].
Fig. 6 shows the TEM images of the parent
sample ZM-2 (Fig. 6a and b). It can be seen that
the individual crystals have an almost rectangular
shape with dimensions of 100800 nm; appropriate
stacking or agglomeration of these crystals result
in the formation of bigger particles with spheroid
shape and dimensions of ~0.52 lm, as also found
by the SEM images (not shown here). The mor-
phology of the crystals/particles observed in the
TEM images of the dealuminated samples was
not changed signicantly compared to the parent
sample, in accord with the high degree of crystal-
linity and microporosity of all the samples. How-
ever, the dierent eect of the degree and the type
of dealumination method on the partial break-
down of the crystals and on the formation of ex-
tra-framework/amorphous phases could be easily
identied. The image shown in Fig. 6c is repre-
sentative of the morphology of the modied crys-
tals/particles observed in the AHFS-dealuminated
sample DZM-7; it can be seen that a small number
of relatively big voids (2050 nm) are formed on
the crystals while the edges of the crystals are not
sharp and well-dened as in the parent sample. In
addition, the phases shown in Fig. 6d have been
observed in DZM-7 and are agglomerates of very
small particles (2040 nm) with irregular shape;
the voids formed between these particles are in
the range of 510 nm. The images of the modied
crystals/particles of the more severely dealumi-
nated, steamed sample DZM-11, shown in Fig. 6
eg, are dierent from those of the AHFS-dealu-
minated sample. The particle shown in Fig. 6e is
more damaged (showing a sponge-like structure)
compared to the one of Fig. 6c, while numerous
holes of dierent sizes (520 nm) can be clearly
seen on the crystals of Fig. 6f. Furthermore, the
phases shown in Fig. 6g are nearly amorphous
(based on electron diraction data), and they
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 381
resemble a wormhole-like structure with pores in
the range of 25 nm; that pore size is in agree-
ment with the mean pore diameter (~2.7 nm) of
DZM-11 which was estimated by the Horvath
Kawazoe method based on the nitrogen adsorp-
tion data.
Fig. 6. Transmission electron micrographs of the parent sample ZM-2 (a,b), the AHSF-dealuminated sample DZM-7 (c,d), and the
steamed sample DZM-11 (eg).
382 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
3.8. Acidity characteristics of the H-ZSM-5 sam-
ples ammonia-temperature programmed desorp-
tion
The number and distribution of acid sites were
tested by ammonia-TPD experiments. By com-
paring the results of previous works regarding
Y-type [23,6163] or ZSM-5 type [6,26,27,6270]
zeolites, it was revealed that ammonia-TPD data
depend largely on the experimental conditions, i.e.
adsorption temperature, ``stripping'' temperature
and time, desorption heating rate, carrier gas ow
rate. In the present study, adsorption of ammonia
took place at 100C in a static system, followed by
stripping with dry pure helium for 1.5 h at the
same temperature. This time was sucient for the
TCD signal to be stabilized in all the tested sam-
ples. Desorbed ammonia was then monitored be-
tween 100C and 700C.
The TPD curves of selected samples are given in
Fig. 7. The TPD curves were deconvoluted into
individual desorption peaks of Gaussian shape, as
also shown in Fig 7. Before applying the decon-
volution procedure, the chromatographs were
baseline corrected by subtracting the ``blank'' TPD
experiments which were performed without the
adsorption of ammonia. The resulting peaks were
classied and grouped into two types of acid
sites with dierent acidity strength, i.e. weak and
strong, according to the temperature of the peak
maximum. The above characterization of the acid
sites is a qualitative indication of how strongly
the ammonia molecules are connected to the acid
sites. A quantitative determination of the acidity
strength of the acid sites from ammonia-TPD data
can be performed by applying appropriate calcu-
lation models [65,71,72]. The temperature range
for the peak maxima of the weak acid sites was set
between 150C and 220C and for the strong acid
sites above ~350C.
The total number of acid sites (mmoles NH
3
per gram hydrated sample), determined for each
sample from the total desorbed ammonia, and the
number of acid sites of weak and strong acidity,
estimated from the relative area of the deconvo-
luted peaks, are given in Table 5. The temperature
of the peak maximum of each individual desorp-
tion step and the ratio of the strong acid sites to
the weak acid sites are also given in Table 5. The
repeatability of the measurements was tested by
performing ve replicates of the TPD measure-
ments for the parent sample ZM-2 keeping all the
Fig. 7. Ammonia-TPD curves of the parent sample ZM-2, the AHSF-dealuminated sample DZM-7, and the steamed samples DZM-8
and DZM-11 (experimental spectra and deconvoluted peaks).
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 383
experimental conditions the same. The resulting
mean value (m) was found to be 1.05 mmoles NH
3
per gram hydrated sample, the standard deviation
(s) was 0.04, and for a condence level of 95%,
1:05 0:05 mmol NH
3
per gram hydrated sample
was calculated. The relative standard deviation
(RSD) was estimated to be 3.6%.
3.9. Eect of the degree and the type of dealumi-
nation method on acidity
In Fig. 8 the total number of acid sites is plotted
against the contents of TAl (Fig. 8a) and of FAl
(Fig. 8b). From Fig. 8a it becomes clear that the
total amount of desorbed ammonia (total number
of acid sites) is linearly correlated with the TAl-
content only for the parent and the dealuminated-
fresh (dealuminated by HCl and AHFS) H-ZSM-5
samples, with the exception of the parent sample
ZM-1. All the hydrothermally dealuminated sam-
ples as well as ZM-1 possess signicant amounts of
EFAl species which apparently have a lower ca-
pability of adsorbing ammonia compared to the
FAl atoms; as a result these samples are below the
line that connects the samples that possess mainly
FAl atoms. On the other hand, when the total
amount of desorbed ammonia is plotted against
Table 5
Acidic characteristics of the H-ZSM-5 samples
Samples Acidic FAl
a
(mmol g
1
)
Number of acid sites (desorbed ammonia) (mmol g
1
sample) Strong/weak
(mole ratio)
Total (0.05) Weak Strong
ZM-1 0.58 1.15 0.56 (218)
b
0.59 (330 436 608) 1.05
ZM-2 0.51 1.05 0.53 (218) 0.52 (342 435 610) 0.98
ZM-3 0.35 0.7 0.33 (208) 0.37 (338 426 593) 1.12
ZM-4 ~0 0.07 0.07 (~150)
DZM-5 0.45 0.89 0.45 (218) 0.44 (339 438 605) 0.98
DZM-6 0.34 0.68 0.35 (208) 0.33 (354 438) 0.94
DZM-7 0.29 0.52 0.25 (199) 0.27 (350 423) 1.08
DZM-8 0.16 0.45 0.27 (151 203) 0.18 (361 508) 0.67
DZM-9 0.07 0.21 0.13 (169 230) 0.08 (360 524) 0.62
DZM-10 0.02 0.11 0.07 (163 221) 0.04 (331 522) 0.57
DZM-11 ~0 0.09 0.09 (160 233)
DZM-12 0.03 0.14 0.10 (156 223) 0.04 (350 524) 0.40
a
Estimated from the FAl content in Table 2 after subtracting the Na

that has not been ion-exchanged by H

.
b
The numbers in parentheses represent the temperature peak maxima at degrees Celsius of the individual/deconvoluted desorption
peaks.
Fig. 8. Correlation of the amount of the total desorbed am-
monia (TPD) with the TAl content (a) and the FAl content (b),
for () parent samples, (_) AHFS dealuminated samples, (M)
HCl dealuminated sample, (j) steamed samples, ( ) steamed/
HCl sample (the dashed line corresponds to a 1:1 analogy).
384 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
the FAl content (Fig. 8b) all the samples fall on
a single line with a high correlation coecient
(0.990). However, the intercept value of this line is
little higher than zero, i.e. 0.08, and the slope value
is 1.82 which means that almost two ammonia
molecules correspond to each FAl atom. It has
been previously shown that the amount of am-
monia of the low-temperature peak is nearly the
same as that of the high-temperature peak in the
TPD spectra of fresh H-ZSM-5 zeolites [27,66,67];
this is now conrmed for hydrothermally dealu-
minated samples as well and is extended down to
very low FAl content values.
In order to examine the relation between the
desorbed ammonia that corresponds to the high-
temperature peaks (strong acid sites) and to the
low-temperature peaks (weak acid sites), the rela-
tive data from Table 5 were plotted against FAl
content, as shown in Fig. 9a and b respectively.
The slope of the mean line in Fig. 9a is 0.99 and
the intercept value is ~0. These results are in
accord with previous works [6,62,63,6669,73],
which suggest that the high-temperature peak of
the TPD spectra or the strong acid sites (and the
related amount of ammonia) correspond to the
FAl atoms of fresh H-ZSM-5 zeolites (OH of
the FAl atoms that give rise to the framework
Br onsted acidity). Based on the data of the present
work, the above suggestion can be extended also to
steam-dealuminated samples with low FAl con-
tent. In addition, the above suggestion of the 1:1
analogy between the ammonia desorbed from
strong acid sites and the FAl content is veried by
the thermal decomposition of the NH
4
ZSM-5
sample; the recorded spectra and the amount of
ammonia detected were similar to those that cor-
respond to the high-temperature peak of the TPD
spectra. Based on the above data and with regard
to the specic experimental conditions of the TPD-
tests, the intercept and the slope estimated from
the equation of the straight line in Fig. 9a can be
used for the determination of the FAl content per
gram sample of the H-ZSM-5 zeolites, irrespective
of the degree of dealumination and the amount of
extra-framework phases present in the samples.
The slope of the mean line for the fresh samples
in Fig. 9b (low-temperature peak) is 1.06 and the
intercept value is ~0; the corresponding values for
the mean line of the steamed samples in Fig. 9b are
1.22 and 0.06. It seems that when only the fresh
samples are considered, the amount of desorbed
ammonia of the low-temperature peak is in a 1:1
analogy with the FAl content, as in the case of the
strong acid sites. However, there is a small amount
of ammonia (~0.07 mmol g
1
sample) that has
been detected in the parent sample ZM-4, which
has practically no acidic FAl; if this small amount
of ammonia, which could be attributed to physical
adsorption and to weak chemisorption on the
terminal SiOH sites, is subtracted from the am-
monia of the low-temperature peak then the above
ratio for the sites that correspond to the low-
temperature peak is somewhat less than 1 (~0.8).
These weak sites become even less after hydro-
thermal treatment (when the sample ZM-4 was
Fig. 9. Correlation of the amount of ammonia desorbed (TPD)
from the (a) strong and the (b) weak acid sites, with the FAl
content, for () parent samples, (_) AHFS dealuminated
samples, (M) HCl dealuminated sample, (j) steamed samples,
( ) steamed/HCl sample (the dashed line corresponds to a 1:1
analogy).
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 385
steamed at 790C the amount of desorbed am-
monia dropped to 0.02 mmol g
1
sample); this can
be explained on the basis of
1
H MAS NMR results
where it has been shown that the SiOH defect
sites in the framework of fresh ZSM-5 zeolites are
``healed'' by the formation of SiOSi bonds after
steaming [74].
The amount of ammonia of the low-tempera-
ture peak is aected by the experimental conditions
of the TPD-tests. When the time of the stripping
step (at 100C) in the TPD-test of the sample ZM-
2 was increased from 1.5 to 24 h, the amount of
ammonia of the low-temperature peak gradually
decreased from 0.53 to ~0.10 mmol g
1
sample,
while the amount of the high-temperature peak
(strong acid sites) remained almost unchanged. It
is thus unclear whether the ammonia of the low-
temperature peak corresponds to actual weak
Br onsted acid sites or to physical adsorption, and
quantication of this peak should done by con-
sidering the experimental conditions of the TPD
experiment as well as the overall characterization
of the dierent types of hydroxyls of the crystalline
or extra-framework phases of the samples. De-
tailed
1
H MAS NMR studies of unsteamed H-
ZSM-5 zeolites with framework Si/Al ~ 1520 did
not reveal the existence of some type of SiOH
which are almost equal in amount with the OH
connected to the FAl atoms [10,7476]. However,
under the TPD-experimental conditions applied in
this study, the molecules of the desorbed ammonia
which are in excess to the FAl content, are shown
to be related to the presence of Al atoms in the
zeolitic framework.
The intercept value of the mean line of the
steamed samples which is little higher than zero
(~0.06 mmol g
1
sample) indicates that these sam-
ples possess some additional sites capable of ad-
sorbing ammonia; these sites can be related to the
Al or SiAl oligomers with low degree of hy-
droxylation, in accord with the
27
Al MAS NMR
data, which have a much lower capability of ad-
sorbing ammonia compared to the FAl atoms, as
discussed above (Fig. 8). Furthermore, this small
amount of ammonia does not increase with the
increase of the EFAl or amorphous phases of the
steamed samples (Table 3) and it is not systemati-
cally related to these species. Related FT-IR stud-
ies on the hydroxyl groups of mildly steamed H-
ZSM-5 zeolites have shown the formation of
weakly acidic hydroxylated EFAl species, which
are no more present under harder steaming con-
ditions [15,16,77]. Furthermore, FT-IR tests on
the adsorption of pyridine on steamed H-ZSM-5
samples have shown that the Lewis acid sites ini-
tially formed become less with increasing severity
of the hydrothermal treatment and they are al-
most eliminated in the most severely steamed sam-
ples [15,77,78]. From the data in Table 5, it can be
seen that the strong/weak acid sites ratio remains
practically unchanged for the parent samples and
those dealuminated by HCl and AHFS (fresh sam-
ples) while it decreases in the steamed samples;
the ratio gets very low in the severely steamed
sample DZM-11 since the amount of ammonia
detected in the TPD spectra of this sample over
350C was negligible. Based on these results, the
eect of steam deactivation of the H-ZSM-5 ad-
ditives in the FCC catalyst on product distribution
can be rationalized in terms of the number of sites
and their acidity strength [1,1820]. It is clear that
the severely steamed samples do not possess any
type of relatively strong acidity (neither Br onsted
nor Lewis) such as that it can be detected by
ammonia-TPD tests and further catalyze the high-
acidity-strength demanding reactions in the crack-
ing of gas-oil. However, the traces of FAl atoms
left in these samples, the SiAl oligomers and to a
lesser extent the terminal SiOH groups may
contribute to reactions like cracking and/or alkene
isomerization which have a signicant eect on
product distribution and properties.
4. Conclusions
The results of the present work showed that the
degree of framework dealumination of ZSM-5 and
the amount/type of the extra-framework phases
formed during dealumination were greatly de-
pended on the method of dealumination and the
severity of the treatment. Although steaming or
steaming/HCl are very eective methods, treatment
with AHFS can be applied for the preparation
of moderately dealuminated (~35%) H-ZSM-5
386 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388
samples with the characteristic of being free of
EFAl.
The eect of the method and the degree of
dealumination on the structural characteristics of
the H-ZSM-5 samples with respect to their relative
crystallinity (XRD) and microporosity (SSA) was
relatively small; no more than 20% of structure
break-down was observed even for severely
steamed and highly dealuminated samples (96%
dealumination). The crystals/particles of the
AHFS dealuminated samples were damaged to
a much lesser degree compared to those of the
steamed samples, as shown by the TEM images,
mainly due to the dierence in the dealumination
process between the two methods. Dierent extra-
framework/amorphous phases with disordered
mesoporosity were observed in the dealuminated
samples; however, the majority of the crystals/
particles retained their integrity to a high degree
in accord with the high crystallinity and micro-
porosity of all the dealuminated samples.
When the ammonia-TPD tests were performed
under certain experimental conditions, the amount
of ammonia that corresponded to the high-tem-
perature peak of the TPD spectra was in a 1:1
analogy with the FAl content of the H-ZSM-5
zeolite samples for a wide range of the FAl content
and irrespective of the dealumination method that
was applied. The extra-framework phases formed
during dealumination of the H-ZSM-5 zeolites had
a very low capability of adsorbing ammonia. The
severely steamed samples showed no strong acid-
ity, rationalizing the performance of the steam-
deactivated H-ZSM-5 zeolites in processes like the
cracking of gas-oil.
References
[1] F.G. Dwyer, T.F. Degnan, in: J.S. Magee, M.M. Mitchell,
Jr. (Eds.), Fluid Catalytic Cracking: Science and Techno-
logy, Studies in Surface Science and Catalysis, vol. 76,
Elsevier, Amsterdam, 1993, p. 499.
[2] N.Y. Chen, W.E. Garwood, Catal. Rev.-Sci. Eng. 28
(1986) 185.
[3] A. Humphries, D.H. Harris, P. O'Connor, in: J.S. Magee,
M.M. Mitchell, Jr. (Eds.), Fluid Catalytic Cracking:
Science and Technology, Studies in Surface Science and
Catalysis, vol. 76, Elsevier, Amsterdam, 1993, p. 41.
[4] W.O. Haag, R.M. Lago, P.B. Weisz, Nature 309 (1984)
589.
[5] P.A. Jacobs, R. von Ballmoos, J. Phys. Chem. 86 (1982)
3050.
[6] N.-Y. Topsoe, K. Pedersen, E.G. Derouane, J. Catal. 70
(1981) 41.
[7] R.M. Lago, W.O. Haag, R.J. Mikovsky, D.H. Olson, S.D.
Hellring, K.D. Schmitt, G.T. Kerr, in: Y. Murakami et al.
(Eds.), New developments in Zeolite Science and Techno-
logy, Studies in Surface Science and Catalysis, vol. 28,
Elsevier, Amsterdam, 1986, p. 677.
[8] Y. Sendoda, Y. Ono, Zeolites 8 (1988) 101.
[9] V.L. Zholobenko, L.M. Kustov, V.B. Kazansky, E.
Loeer, U. Lohse, Ch. Peuker, G. Oehlmann, Zeolites
10 (1990) 304.
[10] E. Brunner, H. Ernst, D. Freude, T. Fr ohlich, M. Hunger,
H. Pfeifer, J. Catal. 127 (1991) 34.
[11] F. Lonyi, J.H. Lunsford, J. Catal. 136 (1992) 566.
[12] A.I. Biaglow, D.J. Parrillo, G.T. Kokotailo, R.J. Gorte,
J. Catal. 148 (1994) 213.
[13] C.S. Triantallidis, N.P. Evmiridis, Ind. Eng. Chem. Res.
39 (9) (2000) 3233.
[14] J. Dakta, S. Marschmeyer, T. Neubauer, J. Meusinger, H.
Papp, F.-W. Sch utze, I. Szpyt, J. Phys. Chem. 100 (1996)
14451.
[15] S.M. Campbell, D.M. Bibby, J.M. Coddington, R.F.
Howe, R.H. Meinhold, J. Catal. 161 (1996) 338.
[16] A. Maijanen, E.G. Derouane, J.B. Nagy, Appl. Surf. Sci.
75 (1994) 204.
[17] N.-Y. Topsoe, F. Joensen, E.G. Derouane, J. Catal. 110
(1988) 404.
[18] J. Scherzer, Catal. Rev.-Sci. Eng. 31 (3) (1989) 215.
[19] J. Biswas, I.E. Maxwell, Appl. Catal. 58 (1990) 1.
[20] C.S. Triantallidis, N.P. Evmiridis, L. Nalbandian, I.A.
Vasalos, Ind. Eng. Chem. Res. 38 (3) (1999) 916.
[21] L. Kubelkova, S. Beran, A. Malecka, V.M. Mastikhin,
Zeolites 9 (1989) 12.
[22] J. Sanz, V. Fornes, A. Corma, J. Chem. Soc., Faraday
Trans. 1 84 (1988) 3113.
[23] C.S. Triantallidis, A.G. Vlessidis, N.P. Evmiridis, Ind.
Eng. Chem. Res. 39 (2) (2000) 307.
[24] M. M uller, G. Harvey, R. Prins, Micropor. Mesopor.
Mater. 34 (2000) 135.
[25] A. De Lucas, P. Canizares, A. Duran, A. Carrero, Appl.
Catal. A: Gen. 154 (1997) 221.
[26] T. Masuda, Y. Fujikata, S.R. Mukai, K. Hashimoto, Appl.
Catal. A: Gen. 172 (1998) 73.
[27] J. Kornatowski, W.H. Baur, G. Pieper, M. Rozwadowski,
W. Schmitz, A. Cichowlas, J. Chem. Soc., Faraday Trans.
88 (1992) 1339.
[28] G. Debras, A. Gourgue, J.B. Nagy, G. De Clippeleir,
Zeolites 6 (1986) 241.
[29] M.W. Anderson, J. Klinowski, J. Chem. Soc., Faraday
Trans. 1 82 (1986) 1449.
[30] S. Namba, A. Inaka, T. Yashima, Chem. Lett. (1984) 817.
[31] P.J. Kooyman, P. van der Waal, H. van Bekkum, Zeolites
18 (1997) 50.
C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388 387
[32] S. Han, D.S. Shihabi, C.D. Chang, J. Catal. 196 (2000)
375.
[33] S. Kumar, A.K. Sinha, S.G. Hegde, S. Sivasanker, J. Mol.
Catal. A: Chem. 154 (2000) 115.
[34] R.J. Argauer, G.R. Landolt, US Patent, 1972, no.
3,702,886.
[35] G. Engelhardt, D. Michel, High-Resolution Solid-State
NMR of Silicates and Zeolites, Wiley, Chichester, 1987,
pp. 205318.
[36] M. St ocker, in: J.C. Jansen et al. (Eds.), Advanced Zeolite
Science and Applications, Studies in Surface Science and
Catalysis, vol. 85, Elsevier, Amsterdam, 1994, p. 429.
[37] C.A. Fyfe, G.C. Gobbi, G.J. Kennedy, J. Phys. Chem. 88
(1984) 3248.
[38] C.A. Fyfe, J.L. Bretherton, L.Y. Lam, J. Am. Chem. Soc.
123 (22) (2001) 5285.
[39] J.B. Nagy, Z. Gabelica, E.G. Derouane, P.A. Jacobs,
Chem. Lett. (1982) 2003.
[40] D. Coster, J.J. Fripiat, Chem. Mater. 5 (1993) 1204.
[41] W. Zhang, X. Bao, X. Guo, X. Wang, Catal. Lett. 60
(1999) 89.
[42] A. Samoson, E. Lippmaa, G. Engelhardt, U. Lohse, H.-G.
Jerschkewitz, Chem. Phys. Lett. 134 (1987) 589.
[43] J. Klinowski, C.A. Fyfe, G.C. Gobbi, J. Chem. Soc.,
Faraday Trans. 1 81 (1985) 3003.
[44] P.J. Grobet, H. Geerts, M. Tielen, J.A. Martens, P.A.
Jacobs, in: H.G. Karge, J. Weitkamp (Eds.), Zeolites as
Catalysts, Sorbents, and Detergent Builders: Applications
and Innovations, Studies in Surface Science and Catalysis,
vol. 46, Elsevier, Amsterdam, 1989, p. 721.
[45] J.M. Cruz, A. Corma, V. Fornes, Appl. Catal. 50 (1989)
287.
[46] L. Kubelkova, V. Seidl, G. Borbely, H.K. Beyer, J. Chem.
Soc., Faraday Trans. 1 84 (1988) 1447.
[47] J.R. Sohn, S.J. DeCanio, J.H. Lunsford, D.J. O'Donnell,
Zeolites 6 (1986) 225.
[48] H. Fichtner-Schmittler, U. Lohse, G. Engelhardt, V.
Patzelova, Cryst. Res. Technol. 19 (1984) K1.
[49] U. M uller, K.K. Unger, in: K.K. Unger et al. (Eds.),
Characterization of Porous Solids, Studies in Surface
Science and Catalysis, vol. 39, Elsevier, Amsterdam,
1988, p. 101.
[50] P. Voogd, J.J.F. Scholten, H. van Bekkum, Colloids Surf.
55 (1991) 163.
[51] B. Sulikowski, J. Klinowski, Appl. Catal. A: Gen. 84
(1992) 141.
[52] J. Kornatowski, M. Rozwadowski, W. Lutz, W.H. Baur,
in: L. Bonneviot, S. Kaliaguine (Eds.), Zeolites: A Rened
Tool for Designing Catalytic Sites, Studies in Surface
Science and Catalysis, vol. 97, Elsevier, Amsterdam, 1995,
p. 259.
[53] P.A. Jacobs, H.K. Beyer, J. Valyon, Zeolites 1 (1981) 161.
[54] O. Terasaki, Y. Sakamoto, J. Yu, Y. Nozue, T. Ohsuna, N.
Ohnishi, Y. Horikawa, K. Hiraga, G. Zhu, S. Qiu, R. Xu,
M. Anderson, Supramol. Sci. 5 (1998) 189.
[55] Y. Sakamoto, M. Kaneda, O. Terasaki, D.Y. Zhao, J.M.
Kim, G. Stucky, H.J. Shin, R. Ryoo, Nature 408 (2000)
449.
[56] A.H. Janssen, A.J. Koster, K.P. de Jong, Angew. Chem.
Int. Ed. 40 (6) (2001) 1102.
[57] C. Choi-Feng, J.B. Hall, B.J. Huggins, R.A. Beyerlein,
J. Catal. 140 (1993) 395.
[58] J. Pellet, D.G. Casey, H.-M. Huang, R.V. Kessler, E.J.
Kuhlman, C.-L. O'Young, R.A. Sawicki, J.R. Ugolini,
J. Catal. 157 (1995) 423.
[59] Z.M.M. Noronha, J.L.F. Monteiro, P. Gelin, Micropor.
Mesopor. Mater. 23 (1998) 331.
[60] R. Dutartre, L.C. de Menorval, F. Di Renzo, D.
McQueen, F. Fajula, P. Schulz, Micropor. Mater. 6
(1996) 311.
[61] L. Forni, E. Magni, E. Ortoleva, R. Monaci, V. Solinas,
J. Catal. 112 (1988) 444.
[62] C.V. Hidalgo, H. Itoh, T. Hattori, M. Niwa, Y. Mura-
kami, J. Catal. 85 (1984) 362.
[63] H.G. Karge, L.C. Jozefowicz, in: J. Weitkamp et al. (Eds.),
Zeolites and Related Microporous Materials: State of the
Art 1994, Studies in Surface Science and Catalysis, vol. 84,
Part A, Elsevier, Amsterdam, 1994, p. 685.
[64] C.J.H. Jacobsen, C. Madsen, T.V.W. Janssens, H.J.
Jakobsen, J. Skibsted, Micropor. Mesopor. Mater. 39
(2000) 393.
[65] L. Forni, F.P. Vatti, E. Ortoleva, Micropor. Mater. 3
(1995) 367.
[66] G.I. Kapustin, T.R. Brueva, A.L. Klyachko, S. Beran, B.
Wichterlova, Appl. Catal. 42 (1988) 239.
[67] W. Reschetilowski, B. Unger, K.-P. Wendlandt, J. Chem.
Soc., Faraday Trans. 1 85 (1989) 2941.
[68] S.B. Sharma, B.L. Meyers, D.T. Chen, J. Miller, J.A.
Dumesic, Appl. Catal. A: Gen. 102 (1993) 253.
[69] V.V. Yushchenko, A.N. Zakharov, B.V. Romanovskii,
Kinet. Catal. 27 (1986) 409.
[70] P.A. Jacobs, J.A. Martens, J. Weitkamp, H.K. Beyer,
Faraday Discuss. Chem. Soc. 72 (1981) 353.
[71] C. Costa, I.P. Dzikh, J.M. Lopes, F. Lemos, F.R. Ribeiro,
J. Mol. Catal. A: Chem. 154 (2000) 193.
[72] T. Masuda, Y. Fujikata, S.R. Mukai, K. Hashimoto, Appl.
Catal. A: Gen. 165 (1997) 57.
[73] R.W. Weber, J.C.Q. Fletcher, K.P. M oller, C.T. O'Con-
nor, Micropor. Mater. 7 (1996) 15.
[74] M. Hunger, Catal. Rev.-Sci. Eng. 39 (1997) 345.
[75] L. Heeribout, P. Batamack, C. Doremieux-Morin, R.
Vincent, J. Fraissard, Colloids Surf. A 115 (1996) 229.
[76] E. Brunner, K. Beck, M. Koch, L. Heeribout, H.G. Karge,
Micropor. Mater. 3 (1995) 395.
[77] J.C. Vedrine, A. Auroux, V. Bolis, P. Dejaifve, C.
Naccache, P. Wierzchowski, E.G. Derouane, J.B. Nagy,
J.-P. Gilson, J.H.C. van Hoo, J.P. van den Berg, J.
Wolthuizen, J. Catal. 59 (1979) 248.
[78] A. Martin, U. Wolf, S. Nowak, B. L ucke, Zeolites 11
(1991) 85.
388 C.S. Triantallidis et al. / Microporous and Mesoporous Materials 47 (2001) 369388

You might also like