Sei sulla pagina 1di 8

Applied Catalysis A: General 364 (2009) 191–198

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Mesoporous ZSM-5 zeolite catalysts prepared by desilication with organic


hydroxides and comparison with NaOH leaching
Sònia Abelló a, Adriana Bonilla a, Javier Pérez-Ramı́rez a,b,*
a
Institute of Chemical Research of Catalonia (ICIQ), Avinguda Paı̈sos Catalans 16, 43007, Tarragona, Spain
b
Catalan Institution for Research and Advanced Studies (ICREA), Passeig Lluı´s Companys 23, 08010, Barcelona, Spain

A R T I C L E I N F O A B S T R A C T

Article history: Hierarchical zeolites combining micro- and mesoporosity were prepared by desilication of ZSM-5 (Si/
Received 27 March 2009 Al = 42) in aqueous solutions of tetraalkylammonium hydroxides (TPAOH, TBAOH). Similarities between
Received in revised form 27 May 2009 the treatment mediated by organic hydroxides and the conventional NaOH leaching comprised (i) the
Accepted 27 May 2009
requirement to work with a starting zeolite in the optimal Si/Al window of 25–50 and (ii) the minor
Available online 6 June 2009
influence of the counter-cation (H+, NH4+, Na+) on the degree of mesoporosity developed. However, both
desilication media presented distinctive kinetic and mechanistic characteristics. The use of organic
Keywords:
hydroxides directly produced the protonic form of the mesoporous zeolite upon calcination, simplifying
Hierarchical zeolites
the final ion exchange with NH4NO3 characteristic of the NaOH treatment. Silicon dissolution in TPAOH
Mesoporous ZSM-5
Desilication (or TBAOH) was much slower than in NaOH, making the demetallation process highly controllable. The
Organic base treatment in the organic base was less selective for silicon extraction, i.e. a higher amount of aluminum
Tetraalkylammonium hydroxide was leached to the solution compared to the NaOH treatment. The differences in porosity development
NaOH and chemical composition of the final solid can be assigned to the effect of the bulky
Benzene alkylation tetraalkylammonium cation. The optimal hierarchical ZSM-5 zeolite (treated in 1 M TPAOH at 338 K
and 8 h) had a mesopore surface area of 160 m2 g 1 and the intrinsic zeolite properties were largely
preserved. This sample displayed improved catalytic performance in the liquid-phase benzene
alkylation with ethylene compared to the parent and NaOH-treated zeolites.
ß 2009 Elsevier B.V. All rights reserved.

1. Introduction typically assigned to the shortened micropore diffusion path


length as a result of a secondary mesopore network of inter- or
Catalytic cracking, alkylation, and isomerization are well- intracrystalline nature.
established processes in industry making use of the unique The selective extraction of framework silicon by treatment in
advantages of zeolites [1]. However, the microporous nature of alkaline solutions, referred to as desilication or base leaching, is
these solid acids imposes diffusion limitations in reactions now a widely used top-down method to prepare hierarchical
involving bulky hydrocarbons. Mass-transfer constraints limit porous zeolites [4–8]. The controlled leaching of Si by OH forms
the catalytic activity and occasionally also the selectivity and intracrystalline mesopores, which facilitate the access and
lifetime. A representative example is the liquid-phase alkylation of diffusion of molecules in the zeolite [9–11]. This modification
benzene with ethylene over conventional ZSM-5 [2], which shows brings important benefits in catalysis associated with higher
poor activity, relatively low ethylbenzene selectivity, and fast activity, selectivity, and/or lifetime [3,11–19]. Strong inorganic
deactivation due to formation of polyethylbenzenes. In such cases, bases (namely NaOH but also KOH and LiOH) have been
it is convenient to increase the accessibility and molecular successfully used to introduce extra mesoporosity in ZSM-5
transport to/from the active sites in zeolites to reach their full [20]. The treatment yields a hierarchical zeolite in the alkaline
catalytic potential. Hierarchical zeolites have emerged as an form due to ion exchange of NH4+ or H+ in the starting zeolite with
important class of materials leading to improved catalytic excess Na+, K+, or Li+ in the alkaline solution. Consequently, an
performance compared to their microporous parents [3]. This is additional ion exchange with NH4NO3 followed by calcination is
required in order to have the mesoporous zeolite in the often
desired protonic form. In addition, using NaOH, most of the
mesoporosity is developed in the first 15 min of the treatment
* Corresponding author at: Institute of Chemical Research of Catalonia (ICIQ),
Avinguda Paı̈sos Catalans 16, 43007, Tarragona, Spain. Tel.: +34 977 920 236;
[5,20]. The fast kinetics of silicon dissolution sometimes offers
fax: +34 977 920 224. limited opportunities to precisely control the process of
E-mail address: jperez@iciq.es (J. Pérez-Ramı́rez). mesopore formation. A recent strategy to tailor the degree of

0926-860X/$ – see front matter ß 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2009.05.055
192 S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198

mesoporosity comprises NaOH treatment of partially detem- 2.2. Characterization methods


plated zeolites [21]. The template-containing regions in the
crystal are far less prone to desilication than template-free The content of silicon and aluminum in the solids and filtrates
regions. Accordingly, the template content basically determines was determined by inductively coupled plasma-optical emission
the volume of attackable zeolite by NaOH and therefore the spectroscopy (ICP-OES) (PerkinElmer Optima 3200RL (radial)).
amount of extra-mesoporosity. Powder X-ray diffraction (XRD) patterns were acquired in a
The use of strong organic bases as the desilicating agent, such Bruker AXS D8 Advance diffractometer equipped with a Cu tube, a
as quaternary ammonium hydroxides, can simplify the final ion- Ge(1 1 1) incident beam monochromator (l = 0.1541 nm), and a
exchange step intrinsic of the inorganic hydroxides. The organic Vantec-1 PSD. Data were recorded in the range of 5–508 2u with an
cation can be decomposed by calcination, rendering the zeolite angular step size of 0.0168 and a counting time of 6 s per step.
in its protonic form. This has been shown by Holm et al. [22], N2 isotherms at 77 K were measured in a Quantachrome
who carried out the one-step desilication and ion exchange of Quadrasorb-SI gas adsorption analyzer. Prior to the measurement,
microporous Na-beta zeolite to mesoporous H-beta zeolite using the samples were degassed in vacuum at 573 K for 12 h. The t-plot
tetramethylammonium hydroxide (TMAOH). These authors method was used to discriminate between micro- and mesopor-
reported that mesopore formation in TMAOH requires longer osity [23]. The mesopore size distribution was obtained by the BJH
times and higher temperatures than in NaOH, and hypothesized model applied to the adsorption branch of the isotherm [24].
that TMA+ is able to stabilize the zeolite structure during Transmission electron microscopy (TEM) was carried out in a
desilication. Surprisingly, no mesopores were formed when the JEOL JEM 1011 microscope operated at 100 kV and equipped with a
starting zeolite was in the protonic form. Besides, the catalytic SIS Megaview III CCD camera. A few drops of the sample suspended
performance of the TMAOH-treated beta was not evaluated and in methanol were placed on a carbon-coated copper grid followed
compared with the parent and NaOH-treated counterparts. by evaporation at ambient conditions.
Deeper insights into the mechanism and kinetics of the Fourier transform infrared spectroscopy (FTIR) was carried out
desilication process mediated by organic hydroxides are required. at 723 K in a Nicolet Magna 550 spectrometer. Prior to the
Herein we have prepared mesoporous H-ZSM-5 by desilication in measurements, the catalysts were pressed in self-supporting discs
aqueous solutions of tetrapropylammonium or tetrabutylammo- (diameter: 1.6 cm, 7 mg cm 2) and pretreated in the IR cell
nium hydroxides. The influence of temperature and time on the attached to a vacuum line at 723 K for 4 h up to 10 6 Torr. Spectra
chemical composition, porosity, morphology, and acidity of the were recorded in the range 1800–4000 cm 1 by co-addition of 64
resulting zeolites has been investigated and complemented by scans and with a nominal resolution of 2 cm 1. All spectra were
catalytic tests in the liquid-phase benzene alkylation. Zeolites with normalized to 10 mg wafers.
different Si/Al ratio (17, 42, and 176) and form (H+, NH4+, Na+) were The heat profiles upon contacting aqueous solutions of TPAOH
used as starting materials. Analogies and differences between (1 M) and NaOH (0.2 M) with the parent zeolite were determined
desilication in tetraalkylammonium hydroxides and sodium in a SuperCRCTM microcalorimeter from Omnicaltech. The zeolite
hydroxide have been drawn. (ca. 100 mg) was introduced in a glass vial sealed with a septum
and located inside a Teflon reactor. The reactor was heated to 338 K
2. Experimental and allowed to stabilize. Thereafter, a volume of the alkaline
solution (3 cm3) was dosed by means of a syringe. Simultaneously,
2.1. Parent zeolites and treatments the same volume of distilled water was added to the reference vial
containing 100 mg of zeolite. Both vials were stirred magnetically
A commercial ZSM-5 zeolite with Si/Al ratio of 42 (CBV 8014, and the heat flow was recorded at intervals of 3 s during 30–
Zeolyst, NH4-form) was primarily used in this study, although 60 min. The enthalpy per gram of zeolite was determined by
ZSM-5 samples with Si/Al ratios of 17 (CBV-3024E, Zeolyst, NH4- integration of the heat flow versus time profiles.
form) and 176 (PZ 2/40, Chemie Uetikon, NH4-form) were also
employed. The as-received zeolites were typically calcined in static 2.3. Alkylation tests
air at 823 K for 5 h using a heating rate of 5 K min 1. These
conditions were identical in subsequent calcinations. Occasionally, The zeolites were tested in the liquid-phase alkylation of
the as-received samples were brought into the sodium form via benzene with ethylene in a 500 cm3 commercial titanium batch
three consecutive exchanges in 0.1 M NaNO3 followed by autoclave (Premex). First, 200 cm3 of benzene were introduced in
calcination. the reactor with approximately 100 mg of powdered catalyst that
Alkaline treatments were carried out in a MultiMax parallel was previously pretreated at 573 K in He for 12 h to remove
reactor system from Mettler Toledo. The reactors were filled with moisture. After purging with N2, the reactor was heated to 438 K
5 cm3 of 1 M aqueous solutions of tetrapropylammonium hydro- under mechanical stirring (1000 rpm). Subsequently, ethylene was
xide (TPAOH) or tetrabutylammonium hydroxide (TBAOH), introduced in the reactor until the molar benzene:ethylene ratio
sealed, and introduced in the reactor block at 338 or 358 K. The was 4:1 and a total pressure of 23 bar was obtained. During the
zeolite powder (166 mg) was added to each reactor and stirred reaction, liquid aliquots (0.2 cm3) were extracted from the reactor,
magnetically at 500 rpm for periods ranging from 30 min to 8 h. which were analyzed with a GC (Agilent 6890N) equipped with a
Afterwards, the zeolite suspension was quenched by immersion of CPSil-8B column and an FID detector.
the reactor in an ice–water mixture and filtered. The resulting
solid was washed with distilled water until pH neutral, dried at 3. Results and discussion
373 K for 12 h, and calcined as described above. For comparative
purposes, the parent zeolite was also treated in sodium hydroxide 3.1. Mesoporosity development: organic versus inorganic hydroxides
using standard conditions (0.2 M NaOH, 338 K, and 30 min) [5],
brought into the ammonium form via three consecutive The results shown in Table 1 and in the illustrations below
exchanges in 0.1 M NH4NO3, and calcined. The notation used correspond to the zeolite with Si/Al = 42 in the protonic form and
along the manuscript was P for the parent zeolite, OT for the the products derived from alkaline treatment. The XRD pattern of
zeolites treated in organic hydroxides, and AT for the reference the parent zeolite (P) exhibits the MFI structure as the only
NaOH-treated zeolite. crystalline phase (Fig. 1). The N2 isotherm presents a high uptake at
S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198 193

Table 1
Chemical composition and textural properties of the parent and alkaline-treated zeolites.

Sample Base C (M) T (K) t (min) Molar Si/Al Sifiltratea (ppm) Alfiltratea Yieldb Vpore Vmicroc Vmesod Smesoc
ratioa (ppm) (%) (cm3 g 1
) (cm3 g 1
) (cm3 g 1
) (m2 g 1)

Solid Filtrate
e
P – – – – 42 – – – – 0.28 0.17 0.11 60
OT-1 TPAOH 1 338 30 40 – – – 95 0.29 0.17 0.12 80
OT-2 TPAOH 1 338 300 38 – – – 90 0.36 0.14 0.22 125
OT-3 TPAOH 1 338 480 38 90 1007 11 82 0.37 0.13 0.24 160
OT-4 TPAOH 1 358 300 32 79 5598 68 61 0.59 0.13 0.46 180
OT-5 TBAOH 1 338 480 – – – – 77 0.38 0.15 0.23 150
AT NaOH 0.2 338 30 26 902 4967 5 62 0.57 0.09 0.48 277
a
ICP-OES.
b
Grams of solid after workup per gram of sample treated.
c
t-plot method.
d
Vmeso = Vpore Vmicro.
e
The parent zeolite from which the various OT and AT samples were derived was CBV 8014 in the protonic form.

low relative pressures, confirming the microporous character volume was unchanged with respect to the parent sample and the
(Fig. 2, left). The micropore volume of 0.17 cm3 g 1 is characteristic mesopore surface area slightly increased to 80 m2 g 1. This is likely
of MFI. The contribution of mesoporosity (Smeso = 60 m2 g 1) is a due to the limited attack by OH to Si–OH defects thereby
consequence of the crystal’s external surface and the surface increasing the roughness of the crystal’s external surface. The yield
roughness. The zeolite was subjected to standard sodium of this treatment was ca. 95%, indicating the very low silicon
hydroxide treatment (AT). The resulting N2 isotherm shows dissolution compared with NaOH. This control experiment enables
enhanced uptake at intermediate pressures (Fig. 2, left), which to conclude the relatively low reactivity of the organic hydroxide
is indicative of the formation of a hierarchical porous system compared to sodium hydroxide, in spite of the equivalent alkalinity
combining micro- and mesoporosity. As shown in Table 1, the of both basic solutions (pH of 1 M TPAOH is 14 and 0.2 M NaOH is
mesopore surface area and mesopore volume increased from 60 to 13.2). Microcalorimetry substantiates the much lower degree of
277 m2 g 1 and from 0.11 to 0.48 cm3 g 1, respectively. Conco- silicon extraction by the organic hydroxide. Fig. 3 shows the heat
mitantly, the micropore volume decreased from 0.17 to flow profiles upon interaction of the zeolite with 0.2 M NaOH or
0.09 cm3 g 1. Mesopore formation is due to the OH -assisted 1 M TPAOH aqueous solutions at 338 K. The total heat released,
selective extraction of framework silicon [7]. Accordingly, the Si/Al determined by integration of the profiles, was ca. 12 times higher
ratio in the resulting solid decreased from 42 in P to 26 in AT with NaOH (81 J g 1) than with TPAOH (7 J g 1). In agreement with
(Table 1). The molar Si/Al ratio in the filtrate amounts to ca. 1000, the limited mesoporosity development concluded from N2
evidencing the high selectivity of NaOH towards silicon leaching. adsorption, the TPAOH profile is mostly associated with adsorption
The yield of the AT treatment, defined as grams of solid after of the TPA+ cations on the zeolite surface with a minor contribution
workup per gram of sample treated was 62%, meaning that ca. 40% of the OH -assisted silicon hydrolysis. The NaOH profile is
of the silicon leached to the solution. However, the hierarchical consistent with the extensive silicon dissolution generated during
zeolite preserves the long-range crystallinity according to X-ray the first 30 min of the treatment.
diffraction (Fig. 1). The slow desilication kinetics of ZSM-5 by tetraalkylammonium
In contrast to NaOH, treatment of H-ZSM-5 in 1 M TPAOH at hydroxides can be advantageous, as it enables a precise control of
338 K for 30 min caused minor alteration of the porosity and the mesopore formation. With this in mind, we screened
chemical composition (sample OT-1 in Table 1). The micropore conditions with the organic hydroxide leading to enhanced silicon
leaching and thereby to substantial mesopore formation. Treat-
ments in 1 M TPAOH were conducted at 338 or 358 K during 300 or
480 min. These relatively long treatments at 338 K effectively
increased the mesopore surface area of the zeolites, reaching 125
and 160 m2 g 1 in OT-2 and OT-3, respectively (Table 1). The N2
isotherms of these samples, e.g. OT-3 in Fig. 2, left, display
enhanced uptake at intermediate pressures, i.e. a typical finger-
print of hierarchical porous zeolites. The values of Smeso and Vmeso
of OT-2 and OT-3 were significantly lower compared to AT.
However, the TPAOH-treated samples offer two benefits: (i) the
better preservation of the micropore volume (25% of Vmicro is lost in
the organic hydroxide in contrast with the 50% loss in NaOH) and
(ii) the higher solid yield (>80%). The above effects are a direct
consequence of the lower proneness of the organic hydroxide to
dissolve framework silicon. It should be stressed that only one
reference NaOH treatment (0.2 M, 338 K, 30 min) was considered
in this work for comparison with the organic hydroxides. Of course,
the use of NaOH at lower concentration (<0.1 M), lower
temperature (<338 K), and shorter time (<15 min) would imply
a higher solid yield, better preserved micropore volume, and a
lower mesopore surface area [5].
The limited Si leaching with the organic hydroxide also brings
Fig. 1. Powder XRD patterns of the parent and alkaline-treated samples. differences in the size of the mesopores. As shown by the BJH pore
194 S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198

Fig. 2. N2 isotherms at 77 K of the parent and alkaline-treated zeolites (left) and the corresponding adsorption BJH pore size distributions (right).

size distribution (Fig. 2, right), the zeolite treated in TPAOH at Another common feature of the treatments in TPAOH and NaOH
338 K shows the formation of smaller mesopores (OT-3, 6–7 nm) is the dependence of the mesopore surface area on the Si/Al ratio of
compared to the NaOH treatment (AT, 10 nm). Holm et al. [22] the parent zeolite. It is well known that a key condition to fabricate
reported the formation of larger pores in H-beta zeolite (covering mesoporous ZSM-5 zeolites by desilication with sodium hydroxide
the broad range of 2–50 nm) by treatment in TMAOH (0.1 M, 323– is to start with samples with a molar Si/Al ratio in the range 25–50
373 K, 2 h), while smaller pores (3–4 nm) were attained by [7]. Treatment of H-ZSM-5 with Si/Al ratios of 17 and 176 using OT-
conventional NaOH treatment. However, the type of zeolite 3 conditions did not significantly alter the mesopore surface area of
framework and starting form, organic base, and experimental the parent samples (from 30–40 m2 g 1 to 60–70 m2 g 1).
conditions differed from those investigated by us. The advantages of the organic base represented by OT-3 vanish
In addition to H-ZSM-5, we conducted desilication experiments if the treatment conditions become too harsh. For instance,
using OT-3 conditions over ZSM-5 in the sodium and ammonium increasing the temperature of the TPAOH treatment to 358 K for
forms. In analogy with NaOH [20], the counter-cation in the 300 min (sample OT-4) induces a somewhat higher N2 uptake at
starting zeolite (H+, Na+, NH4+) had a minor influence on the high p/p0 compared to OT-3 (Fig. 2, left). As a matter of fact, OT-4
mesoporous surface area by silicon extraction with organic resembles the NaOH-treated zeolite judging from the similar
hydroxides. This result contrasts with the limited mesoporosity increase in Vmeso. Besides, the solid yields in OT-4 and AT are
obtained by Holm et al. [22] with TMAOH when the starting beta practically identical (Table 1). Accordingly, under certain condi-
zeolite was in the protonic form, while extensive mesoporosity tions, an equivalent amount of silicon can be leached using organic
was developed with the sodium form. or inorganic hydroxides. Due to the excessive dissolution of silicon
with temperature, the PSD of OT-4 broadens and is shifted to larger
pore sizes (centered at 20 nm, Fig. 2, right). The creation of larger
pores in OT-4 compared to AT explains the lower mesopore surface
area of the former sample. Surprisingly the micropore volume of
OT-4 was decreased to a lesser extent compared to AT (0.13 versus
0.09 cm3 g 1).
The mesoporous zeolites obtained by desilication in organic
hydroxides preserved the long-range crystallographic order, as
shown by X-ray diffraction (Fig. 1). Transmission electron
microscopy confirmed the generation of intracrystalline meso-
porosity in the TPAOH-treated samples (Fig. 4). In good agreement
with the pore size distribution derived from nitrogen adsorption,
OT-3 presents mesopores in the range of 6–10 nm. The extent of
mesopore formation is enhanced in OT-4. The zeolite aggregates
are perforated by the base, resulting in the formation of larger
pores (>20 nm) and wide entrances to the center of the crystal due
to a more pronounced dissolution. The micrographs of the AT and
OT-4 zeolites were rather similar.
Noteworthily, the treatments in NaOH and TPAOH do not only
provoke porosity differences in the resulting materials due to the
distinct Si dissolution kinetics, but also changes in the chemical
composition. The Si/Al ratios of OT-3 and OT-4 are significantly
higher than in AT (Table 1). This is due to the more selective
Fig. 3. Microcalorimetric profiles upon contacting the parent zeolite with (*) 1 M character of the inorganic base towards silicon extraction
TPAOH and (*) 0.2 M NaOH aqueous solutions at 338 K. compared to the organic base. Accordingly, the Si/Al ratio in the
S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198 195

Fig. 4. Transmission electron micrographs of the parent and alkaline-treated zeolites.

filtrate was around 10 times higher in NaOH than in TPAOH, i.e. a mesoporous zeolite is directly obtained by decomposition of the
higher degree of aluminum leaching occurs with the organic organic cation. In contrast, the infrared spectrum of NaOH-treated
hydroxide. zeolite after calcination shows no band at 3610 cm 1 due to Na+
The desilication treatment has been exemplified above for cations occupying ion exchange positions [7]. The Brønsted feature
TPAOH. However, the attainment of substantial mesoporosity
coupled with the preservation of microporosity also applies to
other alkylammonium hydroxides. To check this, treatment in
tetrabutylammonium hydroxide (TBAOH) under the same condi-
tions as OT-3 was carried out (sample OT-5 in Table 1). The
textural parameters of the zeolite (Smeso = 151 m2 g 1 and
Vmicro = 0.15 cm3 g 1) and solid yield largely resemble OT-3. Other
characterization (N2 isotherm and pore size distribution, XRD, TEM)
was equivalent too and is not shown for conciseness. This result
suggests that the nature of the organic cation in the hydroxide (at
least TPA+ and TBA+) does not significantly affect the amount and
type of mesopores. We also conducted treatments with typical Lewis
organic bases, such as triethylamine or triethanolamine. As
expected, the latter did not create mesoporosity in the zeolite. This
is due to the weaker basicity associated with the presence of the lone
pair of electrons on the nitrogen atom, which makes these bases
acting as proton acceptors rather than hydroxide donors.
Fig. 5 shows the FTIR spectra in the OH-stretching region of the
parent and hierarchical zeolites. The contribution at 3610 cm 1
stems from the Si–OH–Al stretching vibration associated with
Brønsted acid sites [25]. This contribution gradually decreases in
OT-3 and OT-4 due to the concomitant loss of micropore volume.
The alkaline treatment simultaneously induces the progressive
development of isolated silanol groups at 3740 cm 1, due to the
enhanced external surface area associated with the formation of
intracrystalline mesoporosity. It should be recalled that the OT-3
and OT-4 samples were obtained by calcination of the TPAOH- Fig. 5. Infrared spectra in the OH stretching region of the parent and alkaline-
treated zeolites, confirming that the protonic form of the treated zeolites.
196 S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198

was regained after three successive ion exchanges with NH4NO3 dissociated into the cation and the hydroxyl anion. In our case,
followed by calcination (AT in Fig. 5). The spectrum of AT shows a whether the cation is Na+, TPA+, or TBA+, the formation of hydrated
weak but detectable band at 3660 cm 1, which can be attributed to ion dipoles or solvent-separated ion pairs can be foreseen.
hydroxyl groups connected to extra-framework Al species [26]. In However, tetraalkylammonium cations are not strongly solvated
agreement, several authors have shown that upon desilication of in aqueous solution compared to Na+, due to their larger effective
ZSM-5 by NaOH, the relative amount of Lewis sites in the cationic diameter [29] (e.g. TPA+ ca. 0.9 nm [30] versus Na+
mesoporous zeolite increases [27,28], pointing to the transforma- 0.19 nm [31]), and the hydrophobicity of their alkyl groups [32],
tion of some of the Brønsted sites into Lewis sites. The 3660 cm 1 causing the attraction of less water molecules. This is confirmed
band cannot be discerned in any of the TPAOH-treated samples. by the lower hydration enthalpy of tetraalkylammonium cations
This is probably connected with a different desilication mechanism (ca. 135 kJ mol 1) with respect to Na+ (406 kJ mol 1), tending to
running with NaOH and the organic hydroxide, which is put make the former cation less soluble [33]. Besides, TPA+ exhibits a
forward in the next section. high structure stabilizing ability, taking into consideration that it
is the standard structure-directing agent in the synthesis of ZSM-
3.2. Desilication mechanism 5. The hydrophobicity of the long hydrocarbon chains on TPA+
motivates propyl groups to interact with the hydrophobic silicate
Our results evidence the slower desilication kinetics in ZSM-5 species rather than with the polar water molecules [34]. Both
by tetraalkylammonium hydroxides compared to NaOH. With the arguments likely make the relatively bulky TPA+ cation being
organic base, longer treatment times and/or higher temperatures more stable on the zeolite surface. Consequently, in the presence
are required to induce substantial intracystalline mesoporosity. of tetraalkylammonium cations, the zeolite surface is better
This feature enables a good control of the pore formation process shielded from the attack by OH with respect to sodium cations
compared to NaOH, as the amount of silicon extracted is (Fig. 6), i.e. the higher affinity of TPA+ for the zeolite combined
importantly reduced. As a consequence, the micropore volume with its intrinsic steric hindrance, slows down the kinetics of the
is preserved to a larger extent, the newly created intracrystalline desilication process.
mesopores are smaller, and a higher solid yield is attained. The In support of this behavior, the presence of TPA+ cations is also
latter parameter can be easily obtained and provides a primary responsible for the formation of smaller mesopores as compared to
measure of the treatment effectiveness based on the weight loss. the treatment in NaOH. This can be explained by the fact that
Besides, higher Si/Al ratios were obtained, i.e. a higher amount of protected regions inhibit extensive silicon extraction. Harsher
aluminum was leached to the solution in TPAOH compared to experimental conditions are able to generate the formation of
NaOH. This is additionally supported by the lower Al content in the larger pores in the organic base-free regions (sample OT-4). By the
resulting OT-3 sample (389 mmol g 1) with respect to AT same reasoning, a higher amount of aluminum was found in the
(533 mmol g 1). These observations reveal the less selective filtrates as TPA+ deposited on the zeolite surface appears to limit
character of the organic hydroxide towards silicon extraction the realumination of the zeolite, contrarily to what it occurs in
compared to the inorganic base. Fig. 6 summarizes the differences NaOH [7,35]. This is in agreement with infrared results, which
observed between both organic and inorganic base treatments. suggest the absence of extra-framework Al species in the TPAOH-
A different demetallation mechanism, mostly induced by the treated samples. The concept of pore growth inhibition with
nature of the organic cation, should be responsible for the above hydroxides of tetralkylammonium cations is somewhat related to
differences. Bearing in mind the equivalent basicity of both the partial detemplation-desilication treatment recently demon-
TPAOH and NaOH solutions, one should consider additional strated over large beta crystals [21]. In the latter case, the
effects related to solvation, stability, and steric hindrance of the quaternary ammonium cation (TEA+) was occluded in the zeolite
species involved during the desilication process. In a highly polar pores and effectively limited the silicon hydrolysis and derived
solvent like water, strong bases like hydroxides are completely mesopore formation.

Fig. 6. Pictorial comparison of zeolite desilication by tetraalkylammonium hydroxides and sodium hydroxide.
S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198 197

3.3. Alkylation performance microporosity and the newly created mesoporosity. The protect-
ing capability of alkylammonium cations enables not only to
Hierarchical zeolites aim at an improved catalytic performance control the extent of mesopore formation but also to design
by combination in a single material of the catalytic power of hierarchical zeolites with flexible properties.
micropores and the improved transport consequence of a The possibly different type of aluminum species in the OT-3 and
complementary mesopore network. The assessment of function- AT samples should not be neglected in the comparison of the
ality of the samples investigated in this work has been procured in ethylbenzene productivity per mol of Al. The realumination
the liquid-phase alkylation of benzene with ethylene. This reaction phenomenon during NaOH leads to additional Lewis sites whereas
suffers from diffusion limitations and the poor transport of these will be mostly absent in the TPAOH-treated zeolites. A
benzene in the microporous crystals enables ethylene molecules systematic study assessing the strength, nature, and location of
to react with the aromatic ring, leading to oligomerization. acid sites in the NaOH and TPAOH-treated zeolites is required to
The production of ethylbenzene (EB) after 400 min of reaction better understand the effect of the distribution of aluminum
over the parent and alkaline-treated samples is shown in Table 2. species on the catalytic performance. This will be tackled in future
The ethylbenzene productivity over the parent zeolite was work.
2 mmol gcat 1. The productivity over the mesoporous AT sample
is 4 times higher with respect to the parent sample. The improved 4. Conclusions
performance over the mesoporous AT sample is attributed to the
enhanced molecular transport to/from the active sites as a Hierarchical zeolites combining micro- and mesoporosity were
consequence of the extensive introduction of mesopores. Zhao successfully prepared by desilication of ZSM-5 (Si/Al = 42) in
et al. [11] have recently reported a doubled conversion in cumene aqueous solutions of tetraalkylammonium hydroxides (TPAOH or
cracking over the mesopore structured ZSM-5 catalyst compared TBAOH). Common features of the treatment with organic hydro-
to the purely microporous counterpart, which was attributed to xides and the conventional NaOH leaching are the requirement to
the 2–3 order of magnitude increased diffusion coefficient. As work with a starting zeolite in the optimal Si/Al window of 25–50
previously reported for mesoporous mordenite [13], the enhanced and the minor influence of the counter-cation (H+, NH4+, Na+) on
selectivity of AT (97%) compared to the parent sample (90%) is also the degree of mesoporosity developed. However, both desilication
consequence of the improved transport of reactants and products media present distinctive kinetic and mechanistic characteristics.
in the shortened micropores of the former sample, which minimize Organic bases are intrinsically less reactive towards silicon
successive alkylation steps (i.e. polyalkylbenzenes formation). dissolution than inorganic hydroxides. This makes the demetalla-
As described in Section 3.2, the use of alkylammonium tion process highly controllable in comparison with the fast silicon
hydroxides as desilicating agents enables to control the extent dissolution kinetics in NaOH. Besides, the use of organic
of mesopore formation while keeping a higher degree of hydroxides directly produced the protonic form of the mesoporous
microporosity than in NaOH. Accordingly, the intrinsic properties zeolite upon calcination, simplifying the final ion exchange with
of the derived samples might induce interesting differences in the NH4NO3 characteristic of the NaOH treatment. The organic
catalytic performance. In fact, the EB productivity of the OT-3 hydroxide is less selective for silicon extraction, i.e. higher Al
sample is 5 times higher than over the parent zeolite, and exceeds leaching compared to NaOH is observed and therefore higher Si/Al
the value of the conventional alkaline-treated sample. The EB ratios in the mesoporous zeolites are attained. The above
productivity corrected by the aluminum content in each catalyst differences in porosity and composition are determined by the
provides the same trend, but the differences are even more nature of the cation in the organic base. In particular, the steric
pronounced. For example, OT-3 is around 4 and 2 times more hindrance induced by the relatively bulky organic cation is
active per Al atom compared to the parent and AT samples, responsible for this distinct demetallation mechanism. The
respectively, indicating the more efficient site utilization. Such TPAOH-treated mesoporous ZSM-5 is an efficient catalyst for the
reactivity in the organic base-treated samples was not necessarily liquid-phase alkylation of benzene with ethylene. For this
expected. In fact, considering the above results, the facilitated particular reaction, our results highlight the importance of tuning
transport through mesopore formation in OT-3 appears not to be mesoporosity formation to facilitate transport and microporosity
the governing factor for explaining the activity differences, as the preservation in hierarchical porous zeolites. The different type of
Smeso of this sample is much lower than that of AT. In contrast, the aluminum species in the TPAOH and NaOH-treated zeolites can
micropore volume in OT-3 is preserved to a larger extent also influence the catalytic performance. Therefore, a more
(0.13 cm3 g 1). systematic characterization of the acidic properties of the samples
In this particular reaction, the intracrystalline small meso- is required in order to properly quantify the eventual effect. Of
pores observed in the OT-3 sample facilitate the access to the course, in terms of practical implementation, a major inconve-
active sites, leading to a more efficient utilization of the crystal nience of using tetraalkylammonium hydroxides as desilication
volume. In parallel, the higher degree of microporosity plays a agent is the high cost compared to NaOH, even though an extra ion-
more significant role, thus increasing the functionality of the exchange step is avoided.
active sites. Accordingly, it is underlined that the improved
catalytic performance of the organic base-treated samples derives
Acknowledgments
from an elegant and precise interplay between the preserved

The Spanish MICINN (CTQ2006-01562/PPQ, PTQ05-01-00980,


Table 2
Catalytic performance of the zeolites in the liquid-phase benzene alkylation with and Consolider-Ingenio 2010, grant CSD2006-0003) and the ICIQ
ethylene. Conditions: Ratio C6H6:C2H4 = 4:1, T = 438 K, P = 23 bar, t = 400 min. Foundation. Dr. Pablo Ballester (ICREA-ICIQ) is acknowledged for
helpful discussions.
Sample Ethylbenzene productivity S(EB) (%)
1 1
mmol EB gcat mmol EB mmol Al References
P 2 7 90
OT-3 10 26 98 [1] H.F. Rase, Handbook of Commercial Catalysts, first edition, CRC Press, Boca Raton,
AT 8 15 97 2000.
[2] A. Corma, V. Martı́nez-Soria, E. Schnoeveld, J. Catal. 192 (2000) 163.
198 S. Abelló et al. / Applied Catalysis A: General 364 (2009) 191–198

[3] J. Pérez-Ramı́rez, C.H. Christensen, K. Egeblad, C.H. Christensen, J.C. Groen, Chem. [19] Y. Li, S. Liu, Z. Zhang, S. Xie, X. Zhu, L. Zhu, Appl. Catal. A 338 (2008) 100.
Soc. Rev. 37 (2008) 2530. [20] J.C. Groen, J.A. Moulijn, J. Pérez-Ramı́rez, Ind. Eng. Chem. Res. 46 (2007) 4193.
[4] M. Ogura, S. Shinomiya, J. Tateno, Y. Nara, M. Nomura, E. Kikuchi, M. Matsukata, [21] J. Pérez-Ramı́rez, S. Abelló, A. Bonilla, J.C. Groen, Adv. Funct. Mater. 19 (2009) 164.
Appl. Catal. A 219 (2001) 33. [22] M.S. Holm, M.K. Hansen, C.H. Christensen, Eur. J. Inorg. Chem. (2009) 1194.
[5] J.C. Groen, L.A.A. Peffer, J.A. Moulijn, J. Pérez-Ramı́rez, Colloids Surf. A 241 (2004) [23] B.C. Lippens, J.H. de Boer, J. Catal. 4 (1965) 319.
53. [24] E.P. Barret, L.G. Joyner, P.H. Halenda, J. Am. Chem. Soc. 73 (1951) 373.
[6] J.C. Groen, L.A.A. Peffer, J.A. Moulijn, J. Pérez-Ramı́rez, Microporous Mesoporous [25] A. Zecchina, S. Bordiga, G. Spoto, D. Scarano, G. Petrini, G. Leofanti, M. Padovan,
Mater. 69 (2004) 29. J. Chem. Soc. Faraday Trans. 88 (1992) 2959.
[7] J.C. Groen, L.A.A. Peffer, J.A. Moulijn, J. Pérez-Ramı́rez, Chem. Eur. J. 11 (2005) [26] I. Kiricsi, C. Flego, G. Pazzuconi, W. O: Parker, R. Millini, C. Perego, G. Bellussi,
4983. J. Phys. Chem. 98 (1994) 4627.
[8] J.C. Groen, J.A. Moulijn, J. Pérez-Ramı́rez, J. Mater. Chem. 16 (2006) 2121. [27] M.S. Holm, S. Svelle, F. Joensen, P. Beato, C.H. Christensen, S. Bordiga, M. Bjørgen,
[9] X. Wei, P.G. Smirniotis, Microporous Mesoporous Mater. 97 (2006) 97. Appl. Catal. A 356 (2009) 23.
[10] J.C. Groen, W. Zhu, S. Brouwer, S.J. Huynink, F. Kapteijn, J.A. Moulijn, J. Pérez- [28] F. Thibault-Starzyk, I. Stan, S. Abelló, A. Bonilla, C. Fernandez, J.-P. Gilson, J. Pérez-
Ramı́rez, J. Am. Chem. Soc. 129 (2007) 355. Ramı́rez, J. Catal. 264 (2009) 11.
[11] L. Zhao, B. Shen, J. Gao, C. Xu, J. Catal. 258 (2008) 228. [29] R.J. Fessenden, J.S. Fessenden, Organic Chemistry, second edition, PWS Publish-
[12] Y. Song, X. Zhu, Y. Song, Q. Wang, Appl. Catal. A 288 (2006) 69. ers, Boston, 1982.
[13] J.C. Groen, T. Sano, J.A. Moulijn, J. Pérez-Ramı́rez, J. Catal. 251 (2007) 21. [30] W.J. Roth, Pol. J. Chem. 80 (2006) 703.
[14] M. Bjørgen, F. Joesen, M.S. Holm, U. Olsbye, K.-P. Lillerud, S. Svelle, Appl. Catal. A [31] F.A. Cotton, G. Wilkinson, Advanced Inorganic Chemistry, fourth edition, John
345 (2008) 43. Wiley & Sons, New York, 1996.
[15] C. Mei, P. Wen, Z. Liu, H. Liu, Y. Wang, W. Yang, Z. Xie, W. Hua, Z. Gao, J. Catal. 258 [32] D.D. Ensor, H.L. Anderson, T.G. Conally, J. Phys. Chem. 78 (1974) 77.
(2008) 243. [33] M.B. McBride, M.M. Mortland, Clays Clay Miner. 21 (1973) 323.
[16] S. Gopalakrishnan, A. Zampieri, W. Schwieger, J. Catal. 260 (2008) 193. [34] M.M. Helmkamp, M.E. Davis, Annu. Rev. Mater. Sci. 25 (1995) 161.
[17] L. Jin, X. Zhou, H. Hu, B. Ma, Catal. Commun. 10 (2008) 336. [35] G. Lietz, K.H. Schnabel, Ch. Peuker, Th. Gross, W. Storek, J. Völter, J. Catal. 148
[18] F. Jin, Y. Cui, Y. Li, Appl. Catal. A 350 (2008) 71. (1994) 562.

Potrebbero piacerti anche