Você está na página 1de 11

A characterization study of the surface acidity of solid

catalysts by temperature programmed methods


Francesco Arena
a,*
, Roberto Dario
b
, Adolfo Parmaliana
a,c
a
Dipartimento di Chimica Industriale, Universita di Messina, Salita Sperone c.p. 29, I-98166, S. Agata, Messina, Italy
b
Industrie Chimiche Caffaro, Direzione Ricerca, P. le Marinotti 1, I-33050 Torviscosa, Udine, Italy
c
Istituto CNR-TAE, Salita S. Lucia 39, I-98126 S. Lucia, Messina, Italy
Received 5 August 1997; received in revised form 10 November 1997; accepted 16 January 1998
Abstract
The surface acidic properties of various solid catalysts (i.e., clay montmorillonite, acid modied smectite clay, HY zeolite,
sulphate-promoted ZrO
2
(SO
2
4
=ZrO
2
) and Cs-exchanged dodecatungstophosphoric acid (H
0.5
Cs
2.5
PW
12
O
40
)) have been
comparatively evaluated by temperature programmed desorption (TPD) measurements of adsorbed NH
3
, C
5
H
5
N and C
6
H
6
.
Mathematical analysis of TPD spectra of adsorbed ammonia indicates the presence of weak, medium and strong acid sites in
all the catalysts. On the basis of the TPD patterns of NH
3
, C
5
H
5
N and C
6
H
6
, the nature (Brnsted and Lewis) of the acid sites
has been ascertained. Catalytic measurements in CH
3
OH dehydration and benzene alkylation indicate that besides the acidic
properties, the diffusional phenomena also play a key role on the reactivity of solid acid catalysts. The superior catalytic
performance of the H
0.5
Cs
2.5
PW
12
O
40
system in the above reactions has been explained in the light of a pseudo-liquid
behaviour. # 1998 Elsevier Science B.V.
Keywords: Solid acid catalysts; Temperature programmed desorption (TPD); Surface acidity; Strength distribution; Benzene
alkylation; Methanol dehydration
1. Introduction
In recent years a considerable interest has been
focused on heterogeneous catalysis of organic reac-
tions by inorganic acid and superacid solids since it
has been found that most heterogeneous catalysed
organic reactions (i.e., cracking, isomerization, alkyl-
ation, etc.) proceed under milder conditions and with
greater selectivity than analogous homogeneous reac-
tions [1,2]. Although the high reactivity, ease of use
and recovery, and low waste and environmental pro-
blems make heterogeneous catalysts very attractive
alternatives to homogeneous acidic reagents, molecu-
lar rearrangements on solid catalysts are in many cases
difcult to control mainly because of strong adsorp-
tion and shape-selectivity phenomena in the porous
structure of the catalyst [3].
Several classes of solid acid and superacid catalysts
have been recently disclosed [2,410]. Namely, it has
been shown that sulphated oxides [57], Cs-modied
phosphotungstic acid (H
0.5
Cs
2.5
PW
12
O
40
) [8,9], pro-
Applied Catalysis A: General 170 (1998) 127137
*Corresponding author.
0926-860X/98/$19.00 # 1998 Elsevier Science B.V. All rights reserved.
PI I S0 9 2 6 - 8 6 0 X( 9 8 ) 0 0 0 4 1 - 6
moted clay montmorillonites [2,4] and exchanged
zeolites [10] exhibit a peculiar acid catalytic beha-
viour. Owing to the inadequacy of conventional titra-
tion techniques [11], problems rise in characterizing
the very strong acidity (H
0
<12) of solid catalysts.
Then, alternative methods based on the study of the
interaction of probe molecules with the catalyst sur-
face by temperature programmed desorption [1125],
calorimetry [13,16,20] and FTIR spectroscopy
[14,17,18,26] have been extensively used. Namely,
due to the intrinsic simplicity of the method, tempera-
ture programmed desorption (TPD) of ammonia [11
13,1521] and pyridine [11,12,1416] result up to date
in the most widely used methods to feature the acidic
properties of solid catalysts providing useful informa-
tion on both the concentration and strength of sites.
Recently, a great interest has also been recorded on the
TPD pattern of aromatic molecules [2225] owing to
the poor basicity of such compounds being able to
detect the strongest acidic sites as well as because of
the potential relevance of such substrates to industrial
alkylation reactions. In this respect, Liu and Hsu [25]
showed that substituted benzenes can be used as
adsorbates to evaluate superacidity of solid catalysts
by the TPD method.
Therefore, in this paper we report a systematic
acidity characterization study of different classes
of solid catalysts by temperature programmed
methods using as adsorbates both conventional
basic probe molecules (e.g., NH
3
and C
5
H
5
N) and
aromatic compounds (e.g., C
6
H
6
). The inuences of
both surface acidity and structural properties on the
reactivity of the various catalysts in methanol
dehydration and benzene alkylation model reactions
are discussed.
2. Experimental
2.1. Materials
v Clay montmorillonite K10 (Sud Chemie product)
and acid modified smectite clay (CDP 92-233,
Laporte product) samples were commercial pro-
ducts.
v HY zeolite was obtained by the thermal decom-
position of the NH
4
Y zeolite precursor further to
calcination at 5008C for 6 h.
v Cs-exchanged dodecaphosphotungstic acid
(CSPW) was prepared by the partial neutralization
of the phosphotungstic acid (H
3
PW
12
O
40
) by an
aqueous solution of Cs
2
CO
3
added drop-wise
according to the procedure elsewhere described
[8,9]. The salt obtained after partial neutralization
of the precursor (H
0.5
Cs
2.5
PW
12
O
40
) formed a
precipitate which was filtered, dried and activated
in a N
2
stream for 2 h at 3008C.
v 3% Sulphate promoted zirconia catalyst (SZ3) was
prepared by impregnating a ZrO
2
sample, pre-
viously obtained by the hydrolysis of zirconium
oxychloride with ammonium hydroxide, with an
aqueous solution of (NH
4
)
2
SO
4
. After impregna-
tion the sample was dried at 1108C for 16 h and
then activated at 6008C for 16 h in N
2
flow.
The list of the samples, along with the relative
notation, BET surface area (SA
BET
) and pore volume
values is presented in Table 1.
Before all measurements, unless otherwise speci-
ed, catalyst samples (0.1 g; d
p
<10
4
cm) were
treated in situ for 4 h in a He ow (25 STP ml/min)
at 4008C, except the CSPW sample which was treated
at 3008C to avoid any thermal decomposition.
2.2. Methods
v Surface area and pore volume measurements were
performed by nitrogen adsorptiondesorption at
1958C and subsequent elaboration of the iso-
therms according to the BET method.
v Ammonia temperature programmed desorption
(NH
3
-TPD) measurements in the range 100
7008C were performed in a conventional flow
apparatus using a linear quartz microreactor
(i.d., 4 mm; l., 200 mm), a heating rate () of
128C/min and a He carrier flow rate of 25 STP
Table 1
List of samples
Notation Sample SA
BET
(m
2
/g)
Pore volume
(cm
3
/g)
K10 Clay montmorillonite 251 0.39
CDP Acid modified
smectite clay
280 0.23
SZ3 3 wt% SO
2
4
/ZrO
2
141 0.18
CSPW H
0.5
Cs
2.5
PW
12
O
40
135 0.17
HY HY zeolite 500 0.41
128 F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137
ml/min. The desorption process was monitored by
a Quadrupole Mass Spectrometer (Thermolab,
Fisons Instruments) connected on line with the
reactor by a fast response (transit time<0.5 s)
heated (1808C) inlet capillary system.
The above apparatus and experimental parameters
(carrier ow rate and ) were used for all TPD
experiments.
After the activation treatment, the samples were
cooled down to 1508C and saturated for 1.5 h in a 5%
NH
3
/He stream (25 STP ml/min). Then, samples were
purged in the He carrier owfor ca. 1 h until a constant
baseline level was attained. The NH
3
-TPD spectra
were obtained from the m/z=15 mass-to-charge signal
ratio.
v Pyridine TPD (Py-TPD) measurements were per-
formed in the range 1507508C after saturation of
activated sample in a 2% C
5
H
5
N/He stream (25
STP ml/min) at 1508Cand a subsequent purging in
He flow at 1808C for 1 h. Py-TPD spectra were
obtained from the m/z=79 mass-to-charge signal
ratio.
v Benzene TPD (Bz-TPD) measurements were per-
formed in the range 203808C after saturation of
the activated sample at room temperature in a 15%
C
6
H
6
/He stream (25 STP ml/min) and a subse-
quent purging (He) at 808C for 1 h [2225]. Bz-
TPD spectra were obtained from the m/z=78
mass-to-charge signal ratio.
v Catalytic measurements in CH
3
OH dehydration
were carried out at 2508C and atmospheric pres-
sure in a pulse mode (V
pulse
, 2 ml) using He as
carrier gas (25 STP ml/min) and a catalyst sample
of 0.02 g. Before catalytic tests the samples were
activated for 1 h at 4008C (3008C for CSPW) in
the He carrier flow. Reaction products were ana-
lysed by a GC using a Poropaq-T column. Under
the adopted conditions the only detected product
was dimethylether.
v Catalytic measurements in liquid-phase alkylation
of benzene with benzylchloride were performed at
808C under nitrogen atmosphere using.0.2 mol of
benzene, 0.01 mol of benzylchloride and a catalyst
sample of 3.3 g previously activated at 2508C for
16 h in a nitrogen flow. Activity data expressed in
terms of the kinetic constant ``k
C7H7Cl
'' were taken
from the plot ``ln[C
7
H
7
Cl]'' vs. reaction time in
the conversion range 3060%.
3. Results
3.1. NH
3
-TPD
The NH
3
-TPD spectra of K10 (a), CDP (b), SZ3 (c),
CSPW (d) and HY (e) catalysts are shown in Fig. 1,
while the temperature of desorption maxima (T
di
) and
the amount of desorbed ammonia are summarized in
Table 2.
K10 (Fig. 1(a)) and CDP (Fig. 1(b)) catalysts are
characterized by an analogous broad desorption pat-
tern, spanned in the range 1506008C and asymmetric
on the high temperature side, with the maximum at
2448C. Also the amount of desorbed NH
3
is similar for
these two samples (250308 mmol/g
cat
) or at least
equal (1.01.1 mmol/m
2
cat
) if compared on a surface
area unit (Table 2).
The spectrum of the SZ3 sample (Fig. 1(c)) exhibits
a main maximum at 2528C, thereafter the rate of
ammonia desorption decreases softly giving rise to
a shoulder with an unresolved maximum at 3978C.
The amount of desorbed ammonia (411 mmol/g) cor-
responds to a specic surface adsorption (2.9 mmol/
m
2
) much larger than that of K10 and CDP catalysts.
Fig. 1. Normalized NH
3
-TPD spectra of (a) K10, (b) CDP, (c) SZ3,
(d) CSPW and (e) HY zeolite.
F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137 129
The CSPW sample exhibits a desorption pattern
(Fig. 1(d)) spanned in the range 1505508C with a
main desorption maximum at 3958C and a less intense
one at lower temperature centred at 2418C accounting
for a rather low amount of desorbed ammonia
(33 mmol/g) also on a surface area basis (0.4 mmol/
m
2
).
The HY zeolite exhibits the highest value of de-
sorbed ammonia (1415 mmol/g), the result being com-
parable with that of the SZ3 sample on a surface area
basis (2.5 mmol/m
2
). The spectrum (Fig. 1(e)) looks
quite symmetric consisting of a broad desorption peak
centred at 3068C and spanned in the T range 150
4508C.
3.2. Py-TPD
The Py-TPD patterns of the studied catalysts are
compared in Fig. 2, while the temperature of desorp-
tion maxima (T
di
) and the amount of desorbed C
5
H
5
N
are summarized in Table 3.
Also the Py-TPD patterns of K10 (Fig. 2(a)) and
CDP (Fig. 2(b)) samples result similar in shape,
temperature of desorption maxima and amount of
desorbed pyridine (see Table 3). Such spectra show
a poorly resolved maximum at 2708C, afterwards the
desorption rate decreases slowly giving rise to a
secondary maximum at 3658C zeroing at ca. 7008C.
The amount desorbed is equal to 60 and 87 mmol/g
Table 2
NH
3
-TPD: Temperature of desorption maxima (T
di
) and amount desorbed
Catalyst T
d1
(8C) T
d2
(8C) Amount desorbed
(mmol/g
cat
),
Amount desorbed
(m
mol
=m
2
cat
)
K10 244 250 1.00
CDP 244 308 1.10
SZ3 252 397 411 2.91
CSPW 241 395 33 0.39
HY 306 1415 2.47
150<T
d1
<3008C; T
d2
>300.
Fig. 2. Normalized Py-TPD spectra of (a) K10, (b) CDP, (c) SZ3,
(d) CSPW and (e) HY zeolite.
Table 3
Py-TPD: Temperature of desorption maxima (T
di
) and amount desorbed
Catalyst T
d1
(8C) T
d2
(8C) T
d3
(8C) Amount desorbed
(mmol/g
cat
)
Amount desorbed
(m
mol
=m
2
cat
)
K10 272 60 0.24
CDP 270 87 0.31
SZ3 267 96 0.68
CSPW 232 320 468 33 0.39
HY 264 555 745 1.14
150<T
d1
<3308C; 300<T
d2
<4508C; T
d3
>450.
130 F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137
for K10 and CDP samples, respectively, resulting
in both cases ca. 25% of the amount of desorbed
ammonia.
The spectrum of the SZ3 system (Fig. 2(c)) is even
broader than that of K10 and CDP samples, outlining a
poorly resolved maximum at 2698C strongly convo-
luted with some other components which give rise to a
shoulder at ca. 4708C, zeroing at ca. 7008C. The
desorbed amount is slightly larger than that of the
K10 and CDP samples, however also in this case
resulting in ca. 25% of the amount of desorbed
ammonia (Table 3).
The CSPW sample shows a spectrum (Fig. 2(d))
characterized by three resolved maxima centred at
2358C, 3208C and 4688C, respectively. The amount of
desorbed pyridine is the lowest (33 mmol/g), and only
in this case it results exactly equal to the amount of
desorbed ammonia (cf. Tables 2 and 3).
The HY zeolite shows a Py-TPD pattern (Fig. 2(e))
characterized by two broad but resolved peaks centred
at 2648C and 5558C, respectively, whose overall
intensity accounts for an amount of desorbed
C
5
H
5
N equal to 745 mmol/g, which corresponds to
ca. 50% of the saturation capacity by ammonia
(Table 2).
3.3. Bz-TPD
The Bz-TPD spectra of K10 (a), CDP (b), SZ3 (c),
CSPW (d) and HY (e) catalysts are comparatively
shown in Fig. 3, while the temperature of desorption
maximum (T
d
) and the amount of desorbed benzene
are summarized in Table 4.
K10 and CDP samples exhibit the lowest T
d
values
(1121168C) being also characterized by the lowest
benzene adsorption capacity (1014 mmol/g). The
peak shape is similar for both catalysts looking rather
narrow with a tail of desorption rendering it asym-
metric on the high T side (Fig. 3(a) and (b)).
The SZ3 (Fig. 3(c)) and CSPW (Fig. 3(d)) systems
show a more symmetric desorption peak characterized
by higher values of T
d
(1491578C). The benzene
adsorption capacity is equal to ca. 65 and 130 mmol/g,
respectively, resulting in more than one order of
magnitude larger than those of K10 and CDP samples
on a surface area unit (see Table 4).
The HY zeolite displays a broad peak characterized
by the highest T
d
value (1718C) which accounts for an
adsorption capacity (ca. 730 mmol/g) quite larger than
those of all the previous systems (Table 4).
3.4. Catalytic dehydration of methanol
The activity data of the various systems in methanol
dehydration are summarized in Table 5 in terms of
percentage CH
3
OH conversion, specic activity
expressed as mol of CH
3
OH converted per gram of
catalyst and an adimensional parameter, F, resulting
from the ratio between the moles of CH
3
OH converted
(per pulse) and the moles of adsorbed ammonia per
unit of mass of catalyst (Table 2).
Fig. 3. Normalized Bz-TPD spectra of (a) K10, (b) CDP, (c) SZ3,
(d) CSPW and (e) HY zeolite.
Table 4
Bz-TPD: Temperature of desorption maximum (T
d
) and amount
desorbed
Catalyst T
d
(8C) Amount desorbed
(mmol/g
cat
)
Amount desorbed
(mmol/m
2
)
K10 112 10 0.04
CDP 116 14 0.05
SZ3 146 66 0.47
CSPW 157 133 1.43
HY 171 738 1.13
F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137 131
K10 and CDP samples show an analogous low level
(6%) of methanol conversion which reects the lowest
specic activity (1.610
4
mol/g) and F values (0.5
0.6). The SZ3 sample provides a slightly higher
methanol conversion (8.5%) and correspondingly a
small rise in specic activity (2.310
4
mol/g) even if
the F value (0.6) still results comparable with that of
the above systems. The CSPWand HY zeolite are the
most active systems showing comparable levels (77
80%) and specic activity values (20.510
4

21.310
4
mol/g). The F value of the HY zeolite
(1.3) is two times larger than those of K10, CDP and
SZ3 samples, while for the CSPWsalt it attains a value
(89.1) more than two orders of magnitude larger than
those calculated for the other systems.
3.5. Liquid-phase alkylation of benzene with
benzylchloride
The activity data of catalysts in the liquid-phase
alkylation of benzene are compared in Table 6 in
terms of kinetic constant of the rate of benzylchloride
conversion (K
C7H7Cl
), whilst selectivity data are not
reported as under the adopted conditions, the fraction
of di-alkylated products was comparable for all the
systems ranging between 10% and 20%.
K10, CDP and HY catalysts show an analogous low
activity as indicated by the similar value of the kinetic
constant (1.610
3
2.610
3
s
1
). The SZ3 catalyst
is much more active, the value of its kinetic constant
(2.110
2
s
1
) being larger by one order of magni-
tude than that of previous systems. At least the CSPW
salt is still the most active catalyst with a value of
k
C7H7Cl
(7.710
2
s
1
) which is ca. four times larger
than that of the SZ3 sample and almost one order of
magnitude larger than those of K10, CDP and HY
zeolite.
4. Discussion
4.1. Mathematical analysis of NH
3
-TPD spectra and
acid strength distribution
It is generally recognized that ammonia is an
excellent probe molecule for testing the acidic proper-
ties of solid catalysts as its strong basicity and small
molecular size allow for detection of acidic sites
located also into very narrow pores [1113,1521].
Generally, IR spectroscopy [17,18], calorimetric
[13,16,20] and TPD [1113,1521] techniques are
employed to achieve information on the interaction
of NH
3
with solid acids. Although there is a wide-
spread use of TPD in the studies of the surface acidity
of solid catalysts, it is worth noting that the NH
3
-TPD
spectra are often poorly resolved and then either
experimental artefacts, such as a change in the activa-
tion and/or saturation treatments, or curve deconvolu-
tion methods are needed to get insights on site
distribution and heat of desorption [1113]. Then,
on the basis of complementary characterization
results, a fairly reliable interpretation of the TPD
pattern of ammonia from solid acid catalysts has also
been attained. Indeed, desorption peaks with maxi-
mum in the range 180250, 280330 and 3805008C
are currently attributed to NH
3
chemisorbed on weak,
medium and strong acid sites, respectively, being yet
not possible to discriminate between Brnsted and
Lewis acidity [1113,1621]. Nevertheless, such stu-
dies prevalently focused on the characterization of
zeolitic materials [1121]; namely, on the basis of a
qualitative comparison of the NH
3
-TPD spectral fea-
tures, the effects of various factors (e.g., chemical
composition, structure, Al/Si ratio, activation tem-
Table 5
Activity data of solid acid catalysts in CH
3
OH dehydration
Catalyst Conversion
(%)
Specific activity
(10
4
mol/g)
F
K10 6.0 1.6 0.6
CDP 6.0 1.6 0.5
SZ3 8.5 2.3 0.6
CSPW 77.0 20.5 89.1
HY 80.0 21.3 1.3
Table 6
Kinetic constant (k
C7H7CI
) of the benzene alkylation reaction on
solid acid catalysts
Catalyst k
C7H7CI
(s
1
)
K10 1.6E3
CDP 2.6E3
SZ3 2.1E2
CSPW 7.7E2
HY 2.1E3
132 F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137
perature, etc.) on the surface acidity of this class of
solids have been assessed [1621]. By contrast, cata-
lysts with different chemical composition and struc-
tural properties (Table 1) are considered in this case;
then a complete rationalization of their NH
3
-TPD
patterns on the basis of peaks shape (Fig. 1) and T
d
values (Table 2) appear rather difcult. In order to
overcome the approximations of a purely qualitative
comparison between the NH
3
-TPD patterns of the
different systems, such experimental curves were
analysed by a mathematical tting program and
deconvoluted into Gauss functions, on the assumption
of a normal distribution of the desorption activation
energies arising from a corresponding heterogeneity
sites distribution [1113]. Namely, it was found that
all the experimental NH
3
-TPD proles (Fig. 1)
resulted fromthe contribution of three Gauss functions
corresponding to three types (i.e., weak, medium and
strong) of sites from which NH
3
is released. For all the
systems this deconvolution analysis provides a very
accurate (r
2
>0.99) tting of the experimental curves as
shown in Fig. 4. Furthermore, according to Karge and
coworkers [11,12] assuming independent rst-order
kinetic for NH
3
desorption from each type of site, n,
with respect to coverage (rate
des.,n
=k
n
; where
k=Aexp(E
d
/RT) and rate
des.,tot
=
n
(rate
des.,n
))
and a constant value of the pre-exponential factor,
A, equal to 2.510
5
s
1
, the distribution of the activa-
tion energies for desorption of NH
3
from the various
sites has been calculated. The distribution curves of
the energy of desorption of all the catalysts are shown
in Fig. 5, while the results of the computational
analysis are summarized in Table 7 in terms of most
probable activation energy of desorption (E
d
), width
of distribution (), percentage population (A) and
concentration of acid sites (C). In spite of the hetero-
geneity of the catalytic systems, the analogous values
of E
d
for the same type of sites in the different
catalysts (Table 7) suggest that under the adopted
experimental conditions the desorption process is
not signicantly affected by intraparticle molecular
diffusion phenomena [1113,16]. In fact, such a
Fig. 4. Deconvolution analysis results of the NH
3
-TPD spectra: (a)
K10, (b) CDP, (c) SZ3, (d) CSPWand (e) HY zeolite. Legend: (&)
experimental data; (- - -) calculated curves for desorption from
three different types of sites; () theoretical curve of overall
desorption.
Fig. 5. Calculated distributions of the activation energy of
desorption of ammonia from three types of sites: (a) K10, (b)
CDP, (c) SZ3, (d) CSPW and (e) HY zeolite.
F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137 133
method was already applied to the analysis of the
NH
3
-TPD and Py-TPD spectra providing information
on desorption activation energies and population of
acidic sites of H-mordenites [11], H-Y [12] and Na-Y
zeolites [13].
Fig. 5 shows for the montmorillonite K10
(Fig. 5(a)) and pillared clay CDP (Fig. 5(b)) samples
the predominance of weak and medium sites, their
overall contribution being ca. 80%, and a relative low
concentration of strong sites in both samples equal to
ca. 20%. The SZ3 catalyst (Fig. 5(c)) bears a relatively
stronger surface acidity [6,7] as the population of
strong sites rises to ca. 35% mostly at the expense
of weak ones whose contribution lowers to 20%.
According to literature data the NH
3
-TPD pattern of
the HY zeolite (Fig. 1(e)) features one prominent
desorption peak [12,13,18] whose maximum value
(3068C) is comparable with that found (2702808C)
by other authors for analogous systems [12] and
considerably shifted to higher T (1001508C) with
respect to that of less acidic Na-exchanged Y zeolites
[13,18]. The deconvolution of this spectrum into three
Gauss functions [12,13] and also their attribution on
the basis of their E
d
values (Table 7) to weak (~77 kJ/
mol), medium (~ 90 kJ/mol) and strong (~103 kJ/
mol) acid sites [1113,1821] is in very good agree-
ment with previous ndings. Then, this system should
be characterized by the lowest percentage of strong
acid sites (ca. 4%) counterbalanced by the largest
concentration (ca. 88%) of medium sites
[12,13,15,18]. The CSPW sample is characterized
by the largest population of strong acid sites (59%)
reecting mostly in a lower concentration of weak (ca.
15%) ones [8,9]. On the basis of these ndings, the
following acidity scale relative to the concentration of
weak, medium and strong sites (Table 7) can be
drawn:
``Weaksites'' CSPWK10<SZ3<CDP ~ HY
``Mediumsites'' CSPWK10<CDP<SZ3HY
``Strongsites'' CSPW<HY ~K10CDP<SZ3
Although the modelling of the NH
3
-TPD spectra
allows for a reliable understanding and comparison
of the acid properties of the catalysts, both in terms of
strength, distribution and concentration of sites, it is
evident that no discrimination between B and L acid
sites can still be made on the basis of NH
3
-TPD
results.
4.2. TPD patterns of various probe molecules and
nature of acid sites
In contrast to ammonia, both experimental results
and theoretical analysis proved that desorption of
pyridine is clearly limited by transport limitations
in microporous solids and for this reason it is not a
suitable probe molecule for the assessment of acid-site
strength and density by the TPD method [16,17].
Indeed, our experimental results signal remarkable
discrepancies, mostly in quantitative terms, in the
evaluation of the surface acidity when pyridine
(Table 3) or ammonia are used as probe molecules
(Table 2). In fact, assuming the same chemisorption
stoichiometry [27] for NH
3
and C
5
H
5
N, with the
exception of the CSPW salt providing the same gure
for pyridine and ammonia desorption (33 mmol/g), for
all the other systems the number of acid sites tested by
pyridine adsorption results much lower than that (20
45%) obtained by NH
3
-TPD data. The reason of such
marked differences must be found both in the different
molecular size of the probe molecules [16,17,20] as
Table 7
Parameters of acidity distribution in solid acid catalysts from analysis of NH
3
-TPD spectra
Sample Weak sites Medium sites Strong sites

E
d
(kJ/mol)
A (%) C
(mmol/g)

E
d
(kJ/mol)
A (%) C
(mmol/g)

E
d
(kJ/mo l)
A (%) C
(mmol/g )
K10 76.0 2.1 30.4 76 88.0 4.1 47.3 118 103.1 4.4 22.3 56
CDP 75.9 2.0 35.4 109 88.0 4.0 46.6 144 103.4 4.2 18.0 55
SZ3 76.0 2.0 20.3 83 87.2 4.3 44.2 182 105.1 4.8 35.5 146
CSPW 76.0 2.0 14.6 5 90.5 3.6 28.1 9 104.0 3.9 58.8 19
HY 76.6 2.0 7.7 109 90.2 5.4 88.6 1254 102.8 5.2 3.7 52
134 F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137
well as in the geometry of pyridine chemisorption
[16,17]. For the CSPW sample, likely possessing an
``open'' structure, the acid sites should be located at
the surface, being accessible at all also to pyridine
molecules. Whereas, the fraction of acid sites located
in the micropores of solids like K10, CDP, SZ3 and
HY zeolite could result partially inaccessible for
pyridine adsorption preventing the full surface satura-
tion. The nature of acid sites can also contribute to
hinder pyridine chemisorption. In fact, while no spe-
cic conguration is required by the formation of the
pyridinium ion on Brnsted sites, coordinated pyri-
dine interacts with its electron pair on Lewis centres in
a conguration perpendicular to the surface sites [17]
implying larger diffusion restrictions in the case of
microporous solids [17]. Furthermore, taking into
account the fact that pyridinium ions tend to decom-
pose readily above 2008C, being less stable than
coordinated pyridine [17], the spectrum of the CSPW
sample essentially monitors a prevalence of Brnsted-
type acidity [8,9], while those of K10, CDP and SZ3
[6,7,28] catalysts signal mostly a Lewis-type acidity.
The spectrum of the HY zeolite the only bearing two
resolved peaks at 2648C and 5558C, could signal a
balanced distribution of Brnsted and Lewis acid sites.
However, the occurrence of strong diffusional phe-
nomena coupled with the peculiarity of the desorption
mechanism of pyridine froma zeolitic matrix enabling
the transformation of pyridinium ion into coordinated
pyridine upon heating at T>4008C [12,17] could
account for the occurrence of the high temperature
peak (Fig. 2).
The modest benzene adsorption capacity of K10
and CDP catalysts (Table 4) strongly supports the
view that such systems possess a similar low concen-
tration of Brnsted acid sites enabling benzene
adsorption via the preferential formation of a -bond-
ing between aromatic ring and surface protons (at
geometry of adsorption) [22,23,27]. Although ben-
zene adsorption of the SZ3 sample is larger than that
of the previous catalysts (Table 4), the concentration
of protonic acid sites in this system represents only a
small fraction (ca. 15%) of the total surface acidity
[6,7,28]. By contrast, the CSPW sample adsorbs an
amount of benzene even larger than the number of acid
sites monitored by NH
3
and C
5
H
5
N adsorption
because of the additional adsorption on Cs

ions
[22,24]. HY zeolite exhibits the largest benzene
adsorption capacity (Table 4) resulting in an amount
equal to that of pyridine adsorption. On this account,
since the same molecular size of benzene and pyridine
should imply analogous molecular restrictions, it is
likely that sites chemisorbing pyridine are the same
involved in benzene adsorption, being essentially of
Brnsted-type [14,22,23,27]. Although the maximum
of benzene desorption for the HY zeolite is the highest
in the series (see Table 4) the broad peak shape
(Fig. 3) along with its poor concentration of strong
acid sites (see Table 7) [14,18] points to the marked
inuence of diffusional effects on the maximum posi-
tion, while, the high T
d
value and the narrower ben-
zene desorption prole of the CSPW sample
undoubtedly prove a stronger surface interaction with
benzene molecules. Even if diffusional effects on SZ3
sample cannot be ruled out, its lower surface area
(Table 1) and the value of the desorption maximum
point to a higher strength of B-sites in these systems
with respect to K10 and CDP catalysts characterized
by the lowest T
d
values (Table 4).
The above characterization ndings allow to shed
light on the reactivity of catalysts in methanol con-
version (Table 5) taking into account that this is a
typical protonic acid-catalysed reaction [29]. In parti-
cular, focusing our attention on the F value which can
be taken as a measure of the specic activity of acid
sites of the various catalysts, it emerges that: (i) the
analogous low F values (0.50.6) of K10, CDP and
SZ3 samples mainly arise from a low concentration
and availability of protonic acid sites; (ii) although the
concentration of B-sites is much larger in HY zeolite,
diffusional restrictions limit the kinetics of the reac-
tion and therefore only a small increase in the F value
(1.3) with respect to previous systems is attained: (iii)
the CSPW salt exhibits a F value (89.1) more than two
orders of magnitude larger than those of the other
systems likely as a consequence of a ``pseudo-liquid''
behaviour enabling a ready availability and mobility
of protons at the surface which generally account for
unusual catalytic properties [8,9,30].
4.3. Reactivity of acid catalysts in benzene alkylation
The marked differences observed in the specic
activity of the studied catalysts in the liquid-phase
alkylation of benzene (Table 6) cannot be rationalized
only on the basis of the acidic properties. In fact, the
F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137 135
formation of the large diphenylmethane molecule
implies remarkable structure effects in HY zeolite
(molecular-shape selectivity) with the consequence
that the reaction could take place at the external
surface of the catalyst grains then resulting in lower
activity (Table 6) [10].
Temperature programmed surface reaction mea-
surements of adsorbed C
7
H
7
Cl (results not reported
here) showed the primary formation of HCl at 80
858C on all the systems and irrespective of their acidic
strength, pointing to the easy activation of ben-
zylchloride likely according to the following reaction
scheme:
C
7
H
7
Cl
(ads)
H

s
HCl
(g)
C
7
H

7
(ads)
:
This nding proves that some other step must be the
rate determining step (r.d.s.) of the heterogeneous
alkylation of benzene. In particular, taking into
account the results of benzene adsorption (Table 4)
it is evident that, with the exception of the HY zeolite
whose reactivity is mostly controlled by its structural
properties [10], stronger is the afnity for the substrate
the more reactive the catalyst is in the alkylation
reaction (Table 6). Such a statement is well supported
by the straight-line relationship between rate of ben-
zene alkylation and amount of adsorbed benzene
(Table 4), shown in Fig. 6. Since the catalytic activa-
tion of benzylchloride into the corresponding carbo-
nation (C
7
H

7
(ads)
) occurs readily on all the systems, the
above ndings likely point to the adsorption of the
substrate as r.d.s. of the liquid-phase alkylation of
benzene over heterogeneous catalysts.
5. Conclusions
The surface acid properties of different solid cata-
lysts have been systematically evaluated by tempera-
ture programmed desorption (TPD) measurements
using various probe molecules.
In particular, the main results of this study can be
summarized as follows:
v TPD of adsorbed ammonia is a reliable method to
feature the strength but not the nature (Brnsted or
Lewis) of surface acid sites in solid acid catalysts;
v mathematical analysis of NH
3
-TPD spectra high-
lights the presence of weak, medium and strong
acid sites on all the catalysts also enabling their
quantitative estimate;
v the comparative evaluation of the TPD patterns of
NH
3
, C
5
H
5
N and C
6
H
6
allows to shed light on the
nature (B or L) of the acidic sites;
v the reactivity in methanol dehydration and ben-
zene alkylation reactions is controlled by both the
acidic and structural properties of solid catalysts;
v a ``pseudo-liquid behaviour'' accounts for the
highest catalytic performance of the
H
0.5
Cs
2.5
PW
12
O
40
salt in both methanol dehydra-
tion and benzene alkylation.
.
Acknowledgements
Dr. S. Hocevar (University of Lubiana, 63100 Slo-
venia) is acknowledged for helpful discussion and
suggestions. The nancial support to this work by
SNIARICERCHE S.c.p.a., via Pomarico, Pisticci
Scalo (Matera) is also gratefully acknowledged.
References
[1] K. Tanabe, in: B.L. Shapiro (Ed.), Heterogeneous Catalysis;
Catalysis by Novel Solid Strong Acids and Superacids, Texas
A&M University Press, College Station, TX, 1984.
[2] F. Figueras, Catal. Rev.-Sci. Eng. 30 (1988) 457.
[3] G.A. Olah, Friedel-Crafts Chemistry, Wiley, New York,
1973.
Fig. 6. Relationships between amount of adsorbed benzene
(Table 4) and rate of benzene alkylation (Table 6) of solid acid
catalysts.
136 F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137
[4] T. Cseri, S. Bekassy, F. Figureras, E. Cseke, L.-C. Menorval,
R. Dutarte, Appl. Catal. 132 (1995) 141.
[5] C. Miao, W. Hua, J. Chen, Z. Gao, Catal. Lett. 37 (1996)
187.
[6] B.H. Davis, R.A. Keogh, R. Srinivasan, Catal. Today 20
(1994) 219.
[7] A. Clearfield, G.P.D. Serrette, A.H. Khazi-Syed, Catal Today
20 (1994) 295 and references therein.
[8] T. Hibi, T. Takahashi, T. Okuhara, M. Misono, Appl. Catal. 73
(1986) 69.
[9] T. Nishimura, T.M. Okuhara, M. Misono, Appl. Catal. 73
(1991) L7 and references therein.
[10] B. Coq, V. Gourves, F. Figueras, Appl. Catal. A 100 (1993)
69.
[11] H.G. Karge, V. Dondur, J. Phys. Chem. 94 (1990) 765.
[12] H.G. Karge, V. Dondur, J. Weitkamp, J. Phys. Chem. 95
(1991) 283.
[13] M.C. Abello, A.P. Velasco, M.F. Gomez, J.B. Rivarola
Langmuir 13 (1997) 2596 and references therein.
[14] R.B. Borade, A. Clearfield, J. Phys. Chem. 96 (1992) 6729.
[15] H. Matsuhashi, H. Motoi, K. Arata, Catal. Lett. 26 (1994)
325.
[16] S.B. Sharma, B.L. Meyers, D.T. Chen, J. Miller, J.A.
Dumesic, Appl. Catal. 102 (1993) 253.
[17] N.-Y. Topse, K. Pedersen, E.G. Derouane, J. Catal. 70
(1981) 41.
[18] C.V. Hidalgo, H. Itoh, T. Hattori, M. Niwa, Y. Hurakami, J.
Catal. 85 (1984) 362.
[19] W. Reschetilowski, B. Unger, K.P. Wendlandt, J. Chem. Soc.,
Faraday Trans. 1 85 (1989) 2941.
[20] G.I. Kapustin, T.R. Brueva, A.L. Klyachko, S. Beran, B.
Wichterlova, Appl. Catal. 42 (1988) 239.
[21] K. Chao, B.-H. Chiou, C.-C. Cho, S.-Y. Jeng, Zeolites 4
(1984) 2.
[22] V.R. Choudary, K.R. Srinivasan, A.P. Singh, Zeolites 10
(1990) 16.
[23] M. Otremba, W. Zajdel, React. Kinet. Catal. Lett. 51 (1993)
473.
[24] M. Otremba, W. Zajdel, React. Kinet. Catal. Lett. 51 (1993)
481.
[25] C.-H. Liu, C.-Y. Hsu, J. Chem. Soc., Chem. Comm. (1992)
1479.
[26] T. Barzetti, E. Selli, D. Moscotti, L. Forni, J. Chem. Soc.,
Faraday Trans. 92 (1996) 1401.
[27] A. Jentis, G. Warecka, J.A. Lercher, J. Mol. Catal. 51 (1989)
309.
[28] C. Morterra, G. Cerrato, C. Emanuel, V. Bolis, in: L. Guczi et
al. (Eds.), New Frontiers in Catalysis, Akademiai Kiado,
Budapest, 1993, p. 2585.
[29] S. Hocevar, J. Lever, J. Catal. 135 (1992) 518.
[30] T. Okuhara, T. Arai, T. Ichiki, K. Younge Lee, J. Mol. Catal.
55 (1989) 293.
F. Arena et al. / Applied Catalysis A: General 170 (1998) 127137 137

Você também pode gostar