You are on page 1of 9

Microporous and Mesoporous Materials 331 (2022) 111640

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Improved accessibility of Na-LTA zeolite catalytic sites for the Knoevenagel


condensation reaction
Juliana Floriano da Silva, Edilene Deise da Silva Ferracine, Dilson Cardoso *
Catalysis Laboratory, Chemical Engineering Department, Federal University of São Carlos, São Carlos, 13565-905, SP, Brazil

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

Keywords: The use of zeolites in basic catalysis is limited when large molecules are involved, such as in the manufacture of
Zeolites fine chemicals, due to difficulty in accessing the micropores of the catalyst. A solution to this problem is to
Mesopores synthesize zeolites containing mesopores that improve the accessibility of the active sites. This work investigates
TPOAC and Activation energy
the synthesis of zeolites 4A according to a bottom-up method employing the organosilane surfactant [3-(tri­
methoxysilyl) propyl] octadecyldimethylammonium chloride (TPOAC), at different concentrations in the syn­
thesis mixture (TPOAC/Al2O3 = 0–0.09), to produce mesopores. The zeolites were evaluated in the Knoevenagel
condensation reaction, which is strongly influenced by the accessibility of the catalytic sites, due to the large
sizes of the molecules involved. Zeolites 4A with mesopores, showing higher external surface areas, were formed
in the presence of TPOAC at the different concentrations tested. Higher conversions were achieved using the
zeolites with mesopores, compared to conventional zeolite 4A, due to improved access of the reactant molecules
to the catalytic sites. In addition, the reaction catalyzed by the zeolite with mesopores had lower activation
energy, compared to use of the conventional zeolite 4A.

1. Introduction been developed, including bottom-up and top-down methods [8]. The
top-down methods, specifically desilicalization and dealuminization
Zeolites are microporous crystalline aluminum silicates that possess procedures [10,14–16], offer simplicity but have several disadvantages,
special properties, including thermal stability, shape selectivity, and including: (i) a reduced quantity of active sites, resulting in low catalytic
high acidity or basicity. Consequently, they are highlighted as an activity; (ii) blocking of the sites by amorphous materials; (iii) partial
important class of industrial catalysts with large-scale applications in amorphization of the zeolite structure; and (iv) being unsuitable for
petroleum refining and the petrochemical industry, as well as having zeolites with high lattice aluminum content, such as zeolite LTA [17,18].
considerable potential in the catalytic synthesis of fine chemicals [1–3]. Desilicalization is usually an appropriate method for zeolites with Si/
Despite their desirable features as catalysts, the presence of only Al ratios in the range 20–50, but is not efficient for the formation of
micropores in the zeolite structure can restrict the diffusion of reactants mesoporosity in zeolites with Si/Al ratios <15. Dealuminization is
or products, reducing access of the molecules to the active sites, which widely used for zeolites with low Si/Al ratios, although this technique
hinders the application of these materials in reactions involving large can result in a substantial loss of structural aluminum, consequently
molecules [3–7]. One way to address the diffusional and accessibility decreasing the quantity of active sites [3].
limitations of zeolites is to create larger pores, such as in the mesopore The bottom-up strategies use a template for the formation of meso­
range [8]. Several studies have reported improvements in access to pores and can be divided into two types, according to the physical state
active sites and diffusion through the channels for zeolites with meso­ of the template, either soluble in the reaction medium (soft templating)
pores, compared to conventional zeolites [9–12]. Cho et al. [13] or insoluble in the reaction medium (hard templating). In soft tem­
observed faster molecular diffusion in zeolites LTA with mesopores, plating, surfactants and organosilanes are used as templates [11–13,
produced using a bottom-up method, compared to exclusively micro­ 19–21], while hard templating employs rigid molds, such as inert carbon
porous zeolite LTA. particles [22]. Such methodologies are suitable for the formation of
Various strategies for the formation of mesopores in zeolites have mesopores in zeolites with any Si/Al ratios.

* Corresponding author.
E-mail address: dilson@ufscar.br (D. Cardoso).

https://doi.org/10.1016/j.micromeso.2021.111640
Received 30 August 2021; Received in revised form 27 November 2021; Accepted 13 December 2021
Available online 16 December 2021
1387-1811/© 2021 Elsevier Inc. All rights reserved.
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

The soft templating synthesis route has received considerable


attention for the formation of mesopores during zeolite synthesis. Zeo­
lites with MFI, LTA, SOD, and FAU structures, containing mesopores,
have been synthesized in the presence of amphiphilic organosilane
surfactants, such as [3-(trimethoxysilyl) propyl] hex­
adecyldimethylammonium chloride (TPHAC). Other organosilanes that
have been used are [3-(trimethoxysilyl) propyl] octadecyldimethy­ Scheme 1. Model Knoevenagel condensation reaction.
lammonium chloride (TPOAC) and [3-(trimethoxysilyl) propyl] dode­
decyldimethylammonium chloride (TPDAC), which mainly differ in the Determination of the average crystallite diameter employed the
carbon chain length of the hydrophobic part. Scherrer equation [26] (Eq. (2)), using the same peaks employed for
Choi et al. [20] reported the synthesis of ZSM-5 and LTA zeolites calculation of the relative intensity.
containing mesopores, using the amphiphilic TPHAC organosilane.
Κ⋅λ
Later, Cho et al. [13] investigated the catalytic activity of an LTA zeolite Dcrist. = (2)
β⋅cos θ
with mesopores in the conversion of methanol to hydrocarbons and
dimethyl ether. The zeolite with mesopores provided high methanol Where: Dcrist. = Average diameter of crystallites;
conversion, high selectivity towards dimethyl ether, and lower deacti­
vation, compared to the conventional zeolite. K = Shape factor for spheres (0.94);
Improvements in the applications of zeolites with mesopores have λ = Wavelength;
mainly been in reactions catalyzed by acidic sites [24], while the use of θ = Diffraction angle;
these zeolites in basic catalysis has been limited to sodalite [11,25]. A β => β2 = FWHM2-FWHM(Si)2
possible reason for the low application of zeolites in basic catalysis is
that the processes usually involve bulky molecules, such as various Where FWHM is the width at half height of the peak referring to the
products in the fine chemical industry [23], which would be too large for 2θ diffraction plane of the sample or Si standard. Isotherms of N2
the zeolite micropores [25]. physisorption at 77 K were obtained using an ASAP 2420 instrument
The present work describes the synthesis of zeolites 4A containing (Micromeritics). The zeolites were firstly treated at 350 ◦ C for 12 h,
mesopores, using the TPOAC organosilane surfactant, and the applica­ under vacuum, to remove water and gases physically adsorbed on the
tion of these zeolites in the model Knoevenagel condensation reaction surfaces. The t-plot method [27] was used to determine the total and
between benzaldehyde and ethyl cyanoacetate. external surface areas, as well as the volumes of micropores and meso­
pores present, selecting the experimental points with linear fits showing
2. Experimental R2 ≥ 0.99. The pore diameter distribution was determined by the NLDFT
method, using the cylindrical geometry pore model (available in the
2.1. Synthesis of the LTA zeolite with mesopores Micromeritics software included with the ASAP 2420 instrument). The
mesopores volume was determined by NLDFT, using the curves for the
The synthesis of the zeolite LTA with mesopores was based on the accumulated volume of pores in the size range 2–50 nm. Isotherms for
procedure described by Cho et al. [13]. The molar composition of the the physisorption of CO2 at 273 K were obtained using an ASAP 2020
synthesis mixture was as follows: 1 Al2O3: 1.5 SiO2: 5.0 Na2O: 298H2O: x instrument (Micromeritics), with the micropores volume of the zeolite
TPOAC, where the concentration of the TPOAC organosilane (x) was LTA being calculated by the Dubinin-Radushkevich method [28], ac­
varied in the range 0–0.09. The zeolites obtained were named as LTA-x. cording to Eq. (3).
A conventional zeolite synthesized in the absence of TPOAC was named
as LTA-0. lnV = (
− RT 2 P0
) × ln2 ( ) + lnVmicro (3)
Firstly, two solutions were prepared: solution A, containing sodium E0 β P
hydroxide (Sigma-Aldrich), sodium metasilicate pentahydrate (Sigma-
where, V is the adsorbed volume, Vmicro is the micropores volume, E0 is
Aldrich), and TPOAC (42%, Sigma-Aldrich); and solution B, containing
the characteristic energy depending on the pore structure, and β is the
sodium hydroxide and sodium aluminate (Sigma-Aldrich). Solution B
affinity coefficient characteristic of the adsorbent.
was poured into solution A and the resulting mixture was stirred for 30
The sizes and morphologies of the particles were evaluated using
min.
scanning electron microscopy (SEM). The global Si/Al ratios of the ze­
In the next step, the reaction mixture was transferred to Teflon cups,
olites were determined by energy dispersive spectroscopy (EDS), with
housed in stainless steel autoclaves, and hydrothermally treated at
the samples dispersed on double sided carbon tapes. These analyses
100 ◦ C for 4 h. The solid formed was recovered by filtration and washed
employed an FEG electron microscope operated at a voltage of 2 kV.
until pH ~8.0. After drying, the samples were calcined by heating for 6 h
Magic angle spinning nuclear magnetic resonance (MAS-NMR)
in a muffle furnace at 550 ◦ C, under an air atmosphere, with a tem­
spectra were obtained using a Bruker AVANCE III spectrometer, oper­
perature ramp of 5 ◦ C/min.
ated at 9.4 T and 79.5 MHz, to determine the 27Al and 29Si chemical
shifts. The 29Si NMR spectra were used to calculate Si/Al for the struc­
2.2. Characterization of the catalysts
tures of the synthesized zeolites [29].
X-ray diffractograms of the samples were obtained using a Rigaku
Miniflex 600 diffractometer operated with Cu Kα radiation, at 40 kV and 2.3. Catalytic tests
15 mA. The goniometer speed was 10◦ (2θ) min− 1 and the scanning
angle (2θ) was between 5◦ and 50◦ . The performances of the zeolites as basic catalysts were evaluated in
The relative intensities (RI) of the diffractograms were obtained the Knoevenagel condensation reaction. The model reaction used was
using Eq. (1), where the reference sample was the conventional zeolite between benzaldehyde and ethyl cyanoacetate (both from Sigma-
(LTA-0). The calculation of RI employed the sum of the areas of the five Aldrich, 99%), with 2-cyano-3-phenylacrylate and water as the prod­
peaks in the 2θ range from 21.5◦ to 34.5◦ . ucts (Scheme 1).
∑ A 20 mL volume of the reaction mixture was prepared, with a 1:1 M
Intensity ​ (zeolite)
RI(%) = ∑ × 100 (1) ratio between the reactants, each at a concentration of 0.56 mol/L, and
Intensity ​ (reference)
ethanol (85% w/w) as the solvent. The mixture was kept at 70 ◦ C, under

2
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

Fig. 1. X-ray diffractograms of the LTA-x zeolites.


Fig. 3. Isotherms for the adsorption and desorption of nitrogen, at − 196 ◦ C, for
the LTA-x zeolites synthesized using different amounts of TPOAC (x = 0–0.09).

reaction and calcining the zeolites after each use, thus regenerating the
catalyst.

3. Results and discussion

3.1. Characterization of the zeolites

Fig. 1 shows the X-ray diffractograms of the conventional zeolite


(LTA-0) and the zeolites synthesized using different amounts of TPOAC
organosilane. Despite the presence of increasing amounts of TPOAC in
the reaction mixture, there was the formation of pure zeolites LTA, since
the five most intense diffractions for this structure ((2 0 0), (2 2 0), (4 4
2), (6 4 2), and (8 2 2)), were observed in all the diffractograms, with no
diffractions of other crystalline phases.
In comparison to the diffractogram for the conventional LTA zeolite
(LTA-0), the intensities of the diffraction peaks for the other zeolites
Fig. 2. Relative intensities of the diffractograms, and diameters of the LTA-x decreased as the concentration of organosilane used in the synthesis
crystallites. increased. The addition of 0.09 mol of TPOAC in the reaction mixture
led to a relative intensity (RI) decrease of around 30% (Fig. 2, line A).
stirring, and the reaction was initiated by the addition of 3% w/w of This could be explained by the decrease of the average crystallite
catalyst. Aliquots were periodically collected, filtered, and analyzed by diameter, since smaller crystals present fewer diffraction planes (Fig. 2,
gas chromatography using a Shimadzu GC-2010 chromatograph line B). It is likely that the diameter decreased because crystallite growth
equipped with a flame ionization detector (FID) and an RTX-1 capillary was hindered by the anchoring of the organosilane on the surface [13].
column. The procedure was performed in duplicate. The isotherm for nitrogen physisorption on the LTA-0 zeolite at
The TOF0 (turnover frequency) at the start of the reaction (t = 0 min) − 196 ◦ C showed that there was negligible nitrogen adsorption, under
was calculated using Eq. (4), which relates the initial conversion rate, these conditions (Fig. 3). According to Breck [30], at this temperature,
(dx/dt)t = 0, to the number of moles of sites per gram of catalyst (molsites). the nitrogen molecules diffuse very slowly in the micropores of this
zeolite, due to the narrow micropores (diameter of 0.42 nm) [30,31].
molBenzaldehyde (dx) .CA0 .VRM Very different isotherm profiles were observed for the zeolites syn­
TOF0 ( ) = dt t=0 (4)
min.molsites molsites .mcatalyst thesized using different amounts of TPOAC (Fig. 3). As the TPOAC
content was increased, there was greater nitrogen adsorption at inter­
The conversion rate at t = 0 min, (dx/dt)t = 0, was determined based
mediate pressures (centered at p/po ~0.7), which could be explained by
on the fit of the irreversible second order model to the kinetic curves.
filling of the mesopores [32]. For these samples, the isotherms were a
Under the conditions employed, the initial benzaldehyde concentration
combination of type I and type IV, indicating the presence of both mi­
(CA0) was 0.56 mol/L, the mass of catalyst (mcatalyst) was 0.0735 g, and
cropores and mesopores.
the reaction volume (VRM) was 20 mL. Given that the chemical formula
The nitrogen physisorption data were used with the t-plot method
of zeolite 4A is Na96Al96Si96O384.216H2O, the molecular mass of the
[27] to evaluate the total external surface areas and the mesopore areas
zeolite is 17520 g/mol. It is known that the number of aluminums in the
(Fig. S1) of the zeolites, with the exception of LTA-0, due to its low ni­
zeolite structure is equivalent to the number of sites (oxygen anions).
trogen adsorption [13]. Estimates were also made of the volumes of the
Finally, given knowledge of the molecular mass and the catalyst mass
micropores (V’micro) and mesopores (V’meso) (Fig. 4).
(mcatalyst), it was possible to determine the number of moles of sites per
Table 1 shows the estimated values for the volumes of the micropores
gram of zeolite (molsites).
and mesopores in the synthesized zeolites, together with the total and
The reuse tests were carried out by removing aliquots after 1 h of

3
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

Fig. 4. Fits obtained using the t-plot method for determination of (a) the micropores volume (V’micro) and (b) the mesopores volume (V’meso) of the LTA-x zeolites.

mesopore surface areas.


Table 1 As shown in Table 1, the zeolites synthesized using greater amounts
Textural properties of the LTA-x zeolites, estimated from the N2 physisorption of TPOAC in the reaction mixture presented higher mesopore volumes
data. and external surface areas. This indicated that the TPOAC acted to form
Zeolite V’micro(a) V’meso(a) Vmeso(b) Stotal(a,c) Sext(a,d) Smeso(e) mesopores, which had a diameter distribution in the range from 5 to 15
(cm3/g) (cm3/g) (cm3/g) (m2/g) (m2/g) (m2/g) nm (Fig. 5).
LTA-0 f
0 0 * * 0 Although nitrogen molecules did not diffuse through the micropores
LTA- 0.0035 0.02 0.02 11 5 6 of the LTA-0 zeolite, the isotherms for the zeolites synthesized in the
0.01
presence of TPOAC (Fig. 3) showed increasing amounts adsorbed at p/
LTA- 0.0038 0.06 0.08 54 29 25
0.04
po < 0.02. This was probably due to the presence of micropores of
LTA- 0.0061 0.13 0.16 99 44 55 greater diameters, caused by defects formed after calcination to remove
0.09 the TPOAC, which allowed the access of nitrogen to the micropore en­
a
Estimated by the t-plot method trances, as shown in Fig. 6. The quantity of micropores (V’micro) accessed
b
Estimated by NLDFT (Fig. 5(b)) by the nitrogen corresponded to the value estimated using the t-plot
c
Total surface area method (Fig. 4(a)), shown in Table 1. In addition, the pore diameter
d
External surface area distribution (Fig. 5) indicated that the defects had sizes between 0.63
e
Mesopores surface area = Stotal-Sext [33]. and 0.83 nm.
f
Method not applicable. Since, under the conditions of the N2 physisorption at − 196 ◦ C, the
nitrogen molecules were unable to access the zeolite 4A micropores,

Fig. 5. (a) Pore diameter distributions and (b) cumulative pore volumes of the zeolites (both determined using the NLDFT method).

4
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

Fig. 6. Defects formed in the zeolite structure after calcination to remove the TPOAC. Source: adapted from Ref. [34].

Fig. 7. (a) Isotherms for the adsorption of CO2 at 273 K by the LTA-x zeolites, and (b) fits obtained using the Dubinin-Radushkevich model.

Table 2
Micropore volumes and chemical compositions of the LTA-x zeolites.
Zeolite Vmicroa (cm3/g) Si/Alb Si/Alc FWHMd Al (%)

LTA-0 0.22 1.02 1.08 ± 0.03 2.8 0.50


LTA-0.01 0.22 1.03 1.10 ± 0.07 3.2 0.49
LTA-0.04 0.21 1.05 1.1 ± 0.1 3.3 0.49
LTA-0.09 0.19 1.10 1.11 ± 0.04 3.9 0.48
a
Dubinin-Radushkevich method.
b 29
Si MAS-NMR.
c
EDS.
d 27
Al MAS-NMR.

equilibrium isotherms for the adsorption of CO2 at 0 ◦ C (Fig. 7(a)) were


used to estimate the micropore volumes of the synthesized zeolites [33].
The micropore volumes (Vmicro) were estimated by the Dubinin-
Radushkevich method (Fig. 7(b)), using CO2 physisorption data, with
the values obtained (Table 2) corresponding to those expected for the
LTA structure [33]. The zeolite synthesized in the presence of the
highest amount of TPOAC (x = 0.09) presented a smaller volume of
micropores. This was probably caused by the decrease of the crystallite
diameter, which led to a greater quantity of defects, due to the greater
contribution of the external area.
In the 29Si NMR spectra (Fig. 8), the chemical shift at δ = − 89.3 ppm
evidenced that the structures of the synthesized zeolites consisted
mainly of Si(OAl)4 units, with silicon atoms linked by an oxygen atom to
four tetrahedral aluminum atoms. There was also a small fraction of Si
(OAlOSi)3 units, indicated by the chemical shift at δ = − 94.5 ppm
[34–36]. A broad signal at δ = − 83.5 ppm, associated with silanol
groups [37,38], was more evident for the zeolites synthesized using
higher amounts of TPOAC, which had higher external surface and
mesopore areas. Deconvolution of the 29Si NMR spectra enabled deter­
mination of the Si/Al ratios in the networks of the synthesized zeolites,
with values close to 1 (Table 2), as expected for zeolite 4A synthesized Fig. 8. 29
Si NMR spectra for the LTA-x zeolites.
under the conditions employed [39].

5
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

and smooth surfaces, which is a typical morphology of zeolite LTA. The


zeolites synthesized using the addition of TPOAC also presented cubic
crystalline structures, but with roughness on the surfaces of the parti­
cles, probably due to movement of the crystallites caused by the
anchoring of TPOAC on their surfaces. Another analyzed feature of the
synthesized zeolites was the average particle diameter and standard
deviation, presented on Table S1. In addition, higher amounts of TPOAC
in the synthesis led to a more homogeneous size distribution for these
materials (Fig. S2).

3.2. Evaluation of the zeolites in the Knoevenagel condensation reaction

Evaluation of the catalytic activity of the zeolites was performed


using the Knoevenagel condensation between benzaldehyde and ethyl
cyanoacetate, producing ethyl 3-phenyl-2-cyanoacrylate and water, as a

27
Fig. 9. Al NMR spectra of the LTA-x zeolites.

The 27Al MAS-NMR spectra (Fig. 9) showed the presence of a single


signal centered at δ = 58 ppm, corresponding to tetra-coordinated
aluminum species in the zeolite structure (AlV). Broadening of this
signal in the spectra for the zeolites synthesized using higher amounts of
TPOAC (Table 2) was probably due to distortions in the symmetry of the
bonds of Al atoms [40], demonstrating that these structures had lower
organization, as also indicated by the lower relative intensities of the
X-ray diffractograms. The signal centered at δ = 0 ppm, corresponding to
octahedral aluminum species, was not present, suggesting that all the
aluminum atoms existed within the zeolite lattice. This possibility was
also supported by the Si/Al ratio, determined by 29Si MAS-NMR, of
approximately 1.0, as well as by the global Si/Al ratio (Table 2).
Fig. 10(a)-(d) show micrographs of the zeolites synthesized using Fig. 11. Conversions of benzaldehyde obtained using the LTA-x zeolites in the
different quantities of organosilane. The conventional LTA-0 zeolite Knoevenagel condensation reaction at 70 ◦ C, with fitting using the irreversible
(Fig. 10(a)) exhibited cubic crystalline structures with truncated edges second order model (continuous lines).

Fig. 10. Micrographs of the LTA-x zeolites: (a) x = 0; (b) x = 0.01; (c) x = 0.04; (d) x = 0.09.

6
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

Table 3
Performances of the LTA-x zeolites used in the Knoevenagel condensation re­
action at 70 ◦ C.
Zeolite k2 (L. mol− 1. R2 (dx/dt)0 (10− 2. TOF0
min− 1) min− 1) (min− 1)

LTA-0 0.0107 0.995 0.60 0.17


LTA- 0.0105 0.999 0.59 0.16
0.01
LTA- 0.0181 0.995 1.0 0.28
0.04
LTA- 0.0280 0.991 1.6 0.44
0.09

Table 4
Activation energies for the Knoevenagel condensation reactions catalyzed by the
LTA-x zeolites.
Zeolite Ea (kJ.mol− 1) R2

LTA-0 40.9 0.952


LTA-0.09 33.6 0.971

Fig. 13. Evaluation of catalytic stability of LTA-0.09 catalyst.

identical for all the zeolites (Table 2), the number of basic sites also
remained almost constant for the zeolites synthesized using different
quantities of TPOAC. In turn, the strength of the anionic sites is mainly
dependent on the acidity of the charge-compensating cation and the
aluminum content in the zeolite network [41,42]. Since the
charge-compensating cation (Na+) was the same and the aluminum
content was almost constant, it could be assumed that the strength of the
basic sites remained unchanged. Therefore, the higher conversions
shown in Fig. 11 were due to the improved access of reactant molecules
to the active sites. This improvement was due to the increasing presence
of mesopores and higher external surface areas of the zeolites, caused by
use of the TPOAC organosilane in the synthesis, as shown in Table 1.
For determination of the activation energies of the Knoevenagel re­
actions in the presence of the LTA-0 and LTA-0.09 zeolites, the kinetic
constants were firstly obtained at different temperatures, using the fits of
the irreversible second order kinetic model, as also described by
Ref. [43] (Fig. S3 and Table S3). These values were then used to obtain
the activation energies from the linearized Arrhenius equation plots
Fig. 12. Arrhenius plots used to obtain the activation energies for the LTA-
(Fig. 12).
0 and LTA-0.09 zeolites.
The activation energy for the reaction catalyzed by the zeolite con­
taining mesopores (LTA-0.09) was lower than for the reaction catalyzed
model reaction (Scheme 1). This reaction is highly influenced by the
by the conventional zeolite (LTA-0). This decrease could be explained by
accessibility of the catalytic sites, due to the large sizes of the molecules
the presence of mesopores, which decreased the activation energy
involved.
related to diffusion [8,44], consequently reducing the reaction activa­
Increase of the quantity of TPOAC used in synthesis of the zeolites
tion energy. Therefore, the presence of the mesopores reduced the
resulted in higher conversions (Fig. 11), as also evidenced by the kinetic
diffusional restrictions of the molecules, facilitating access to the cata­
constants (Table 3) obtained from the fits of the irreversible second
lytic sites and lowering the energy barriers that needed to be overcome
order model (Equation S(1)). Table 3 also shows the conversion rates,
for the reactions to occur.
which were used to obtain the TOF0 values. Substantial increases were
After 3 consecutive uses (Fig. 13), about 90% of the initial conver­
observed for the zeolites synthesized using higher amounts of TPOAC,
sion remains. Furthermore, it was verified that the structure of zeolite
with TOF0 for the LTA-0.09 zeolite being around 2.5-fold higher than
LTA-0.09 was maintained after its reuse (Fig. S4), showing that this
TOF0 for the conventional LTA-0 zeolite. Note that LTA-0 and LTA-0.01
catalyst is stable and promising for applications in heterogeneous basic
zeolites have similar TOF0 values, probably because the amount of
catalysis, since it can be recovered and reused.
mesopores formed in LTA-0.01 was small (Smeso = 6 m2) when compared
to LTA-0.04 and LTA-0.09 zeolites (Table 1), this amount being insuf­
4. Conclusions
ficient to increase the catalytic activity (see Table 4).
The higher TOF0 values for the zeolites synthesized using higher
Type 4A zeolites containing mesopores were successfully synthesized
amounts of TPOAC could have been due to three factors, namely the
using the TPOAC organosilane surfactant to form mesopores. The crys­
quantity, strength, and accessibility of the basic sites. The basic sites of
tallite size and the textural properties of these zeolites could be
zeolites are associated with oxygen anions in the lattice, which are
controlled by adjusting the amount of TPOAC used in the reaction
generated by the insertion of aluminum atoms in the zeolite structure, so
mixture. The zeolites containing mesopores provided higher conversions
the quantity of basic sites is the same as the number of structural
of benzaldehyde in the Knoevenagel condensation reaction, compared to
aluminum atoms [41,42]. Since the lattice Al content was virtually
a conventional zeolite 4A (LTA-0). This was due to the increased access

7
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

of molecules to the active sites, as a result of the higher external surface [12] T. Yutthalekha, C. Wattanakit, C. Warakulwit, W. Wannapakdee,
K. Rodponthukwaji, T. Witoon, J. Limtrakul, Hierarchical FAU-type zeolite
area and mesoporosity.
nanosheets as green and sustainable catalysts for benzylation of toluene, J. Clean.
Prod. 142 (2017) 1244–1251, https://doi.org/10.1016/j.jclepro.2016.08.001.
CRediT authorship contribution statement [13] K. Cho, H.S. Cho, L.C. De Ménorval, R. Ryoo, Generation of mesoporosity in LTA
zeolites by organosilane surfactant for rapid molecular transport in catalytic
application, Chem. Mater. 21 (2009) 5664–5673, https://doi.org/10.1021/
Juliana Floriano da Silva: Conceptualization, Methodology, Vali­ cm902861y.
dation, Formal analysis, Investigation, Data curation, Writing, Writing – [14] M.C. Silaghi, C. Chizallet, J. Sauer, P. Raybaud, Dealumination mechanisms of
zeolites and extra-framework aluminum confinement, J. Catal. 339 (2016)
review & editing, Visualization. Edilene Deise da Silva Ferracine:
242–255, https://doi.org/10.1016/j.jcat.2016.04.021.
Conceptualization, Writing, Writing – review & editing. Dilson Car­ [15] J. Van Aelst, D. Verboekend, A. Philippaerts, N. Nuttens, M. Kurttepeli,
doso: Conceptualization, Methodology, Investigation, Validation, E. Gobechiya, M. Haouas, S.P. Sree, J.F.M. Denayer, J.A. Martens, C.E.
Formal analysis, Resources, Data curation, Writing, Writing – review & A. Kirschhock, F. Taulelle, S. Bals, G.V. Baron, P.A. Jacobs, B.F. Sels, Catalyst
design by NH4OH treatment of USY zeolite, Adv. Funct. Mater. 25 (2015)
editing, Visualization, Supervision, Project administration, Funding 7130–7144, https://doi.org/10.1002/adfm.201502772.
acquisition. [16] D. Verboekend, G. Vilé, J. Pérez-Ramírez, Mesopore formation in usy and beta
zeolites by base leaching: selection criteria and optimization of pore-directing
agents, Cryst. Growth Des. 12 (2012) 3123–3132, https://doi.org/10.1021/
Declaration of competing interest cg3003228.
[17] R. Chal, C. Gérardin, M. Bulut, S. VanDonk, Overview and industrial assessment of
synthesis strategies towards zeolites with mesopores, ChemCatChem 3 (2011)
The authors declare that they have no known competing financial 67–81, https://doi.org/10.1002/cctc.201000158.
interests or personal relationships that could have appeared to influence [18] J. Vernimmen, V. Meynen, P. Cool, Synthesis and catalytic applications of
the work reported in this paper. combined zeolitic/mesoporous materials, Beilstein J. Nanotechnol. 2 (2011)
785–801, https://doi.org/10.3762/bjnano.2.87.
[19] K. Cho, R. Ryoo, S. Asahina, C. Xiao, M. Klingstedt, A. Umemura, M.W. Anderson,
Acknowledgments O. Terasaki, Mesopore generation by organosilane surfactant during LTA zeolite
crystallization, investigated by high-resolution SEM and Monte Carlo simulation,
Solid State Sci. 13 (2011) 750–756, https://doi.org/10.1016/j.
The authors are grateful to CNPq for the provision of financial sup­ solidstatesciences.2010.04.022.
port (grant #141843/2020-9), and to the Structural Characterization [20] M. Choi, H.S. Cho, R. Srivastava, C. Venkatesan, D.H. Choi, R. Ryoo, Amphiphilic
Laboratory (LCE/DEMa) of UFSCar, and CAPES (Coordenação de organosilane-directed synthesis of crystalline zeolite with tunable mesoporosity,
Nat. Mater. 5 (2006) 718–723, https://doi.org/10.1038/nmat1705.
Aperfeiçoamento de Pessoal de Nível Superior – Brasil) #001 for [21] A. Inayat, I. Knoke, E. Spiecker, W. Schwieger, Assemblies of mesoporous FAU-type
financial support and the Laboratory of Structural Characterization. The zeolite nanosheets, Angew. Chem. Int. Ed. 51 (2012) 1962–1965, https://doi.org/
authors thank the Nanotechnology National Laboratory for Agriculture 10.1002/anie.201105738.
[22] C.J.H. Jacobsen, C. Madsen, J. Houzvicka, I. Schmidt, A. Carlsson, Mesoporous
(LNNA/Embrapa Instrumentation/FAPESP #2018/01258-5) for the zeolite single crystals [2], J. Am. Chem. Soc. 122 (2000) 7116–7117, https://doi.
NMR analyses. org/10.1021/ja000744c.
[23] M.S. Holm, E. Taarning, K. Egeblad, C.H. Christensen, Catalysis with hierarchical
zeolites, Catal. Today 168 (2011) 3–16, https://doi.org/10.1016/j.
Appendix A. Supplementary data cattod.2011.01.007.
[24] M.F. Othman, S.S. Lapari, Z. Ramli, S. Triwahyono, Sintesis sodalit mesoliang
Supplementary data to this article can be found online at https://doi. dengan campuran templat-templat kation amonium kuaterner untuk tindak balas
kondensasi knoevenagel, Malaysian J. Anal. Sci. 21 (2017) 860–870, https://doi.
org/10.1016/j.micromeso.2021.111640.
org/10.17576/mjas-2017-2104-12.
[25] D. Verboekend, T.C. Keller, S. Mitchell, J. Pérez-Ramírez, Hierarchical FAU- and
References LTA-type zeolites by post-synthetic design: a new generation of highly efficient
base catalysts, Adv. Funct. Mater. 23 (2013) 1923–1934, https://doi.org/10.1002/
adfm.201202320.
[1] J. Cejka, G. Centi, J. Perez-Pariente, W.J. Roth, Zeolite-based materials for novel
[26] A.L. Patterson, The scherrer formula for X-ray particle size determination, Phys.
catalytic applications: opportunities, perspectives and open problems, Catal. Today
Rev. 56 (1939) 978–982, https://doi.org/10.1103/PhysRev.56.978.
179 (2012) 2–15, https://doi.org/10.1016/j.cattod.2011.10.006.
[27] J.H. de Boer, B.G. Linsen, T. van der Plas, G.J. Zondervan, Studies on pore systems
[2] G. Giannetto, Zeolitas: Caracteristicas, Propriedades Y Aplicaciones Industriales,
in catalysts. VII. Description of the pore dimensions of carbon blacks by the t
Editorial Innovación Tecnológica, Caracas, 1990.
method, J. Catal. 4 (1965) 649–653, https://doi.org/10.1016/0021-9517(65)
[3] J. Pérez-Ramírez, C.H. Christensen, K. Egeblad, C.H. Christensen, J.C. Groen,
90264-2.
Hierarchical zeolites: enhanced utilisation of microporous crystals in catalysis by
[28] M.M. Dubinin, The potential theory of adsorption of gases and vapors for
advances in materials design, Chem. Soc. Rev. 37 (2008) 2530–2542, https://doi.
adsorbents with energetically nonuniform surfaces, Chem. Rev. 60 (1960)
org/10.1039/b809030k.
235–241, https://doi.org/10.1021/cr60204a006.
[4] A. Sachse, C. Wuttke, U. Díaz, M.O. De Souza, Mesoporous y zeolite through ionic
[29] Michael Stöcker, Product characterization by NMR, n.d. http://www.iza-online.
liquid based surfactant templating, Microporous Mesoporous Mater. 217 (2015)
org/synthesis/VS_2ndEd/NMR.htm.
81–86, https://doi.org/10.1016/j.micromeso.2015.05.049.
[30] D.W. Breck, Crystalline molecular sieves, J. Chem. Educ. 41 (1964) 678–689,
[5] J. García-Martínez, M. Johnson, J. Valla, K. Li, J.Y. Ying, Mesostructured zeolite y -
https://doi.org/10.1021/ed041p678.
high hydrothermal stability and superior FCC catalytic performance, Catal. Sci.
[31] M. Thommes, Chapter 15 Textural Characterization of Zeolites and Ordered
Technol. 2 (2012) 987–994, https://doi.org/10.1039/c2cy00309k.
Mesoporous Materials by Physical Adsorption, Elsevier B.V., 2007, https://doi.org/
[6] R. Chal, T. Cacciaguerra, S. Van Donk, C. Gérardin, Pseudomorphic synthesis of
10.1016/S0167-2991(07)80803-2.
mesoporous zeolite Y crystals, Chem. Commun. 46 (2010) 7840–7842, https://doi.
[32] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol,
org/10.1039/c0cc02073g.
T. Siemieniewska, Reporting physisorption data for gas/solid systems with special
[7] D. Verboekend, N. Nuttens, R. Locus, J. Van Aelst, P. Verolme, J.C. Groen, J. Pérez-
reference to the determination of surface area and porosity, Pure Appl. Chem. 57
Ramírez, B.F. Sels, Synthesis, characterisation, and catalytic evaluation of
(1985) 603–619, https://doi.org/10.1351/pac198557040603.
hierarchical faujasite zeolites: Milestones, challenges, and future directions, Chem.
[33] J. García-Martínez, D. Cazorla-Amorós, A. Linares-Solano, Further evidences of the
Soc. Rev. 45 (2016) 3331–3352, https://doi.org/10.1039/c5cs00520e.
usefulness of CO2 adsorption to characterise microporous solids, Stud. Surf. Sci.
[8] K. Li, J. Valla, J. Garcia-Martinez, Realizing the commercial potential of
Catal. 128 (2000) 485–494, https://doi.org/10.1016/s0167-2991(00)80054-3.
hierarchical zeolites: new opportunities in catalytic cracking, ChemCatChem 6
[34] M.D.G. Engelhardt, High-resolution Solid-State NMR of Silicates and Zeolites, John
(2014) 46–66, https://doi.org/10.1002/cctc.201300345.
Wiley and Sons, Chichester, 1987.
[9] C.H. Christensen, K. Johannsen, E. Törnqvist, I. Schmidt, H. Topsøe, C.
[35] A.G. Stepanov, Basics of Solid-State NMR for Application in Zeolite Science:
H. Christensen, Mesoporous zeolite single crystal catalysts: diffusion and catalysis
Material and Reaction Characterization, Elsevier B.V., 2016, https://doi.org/
in hierarchical zeolites, Catal. Today 128 (2007) 117–122, https://doi.org/
10.1016/B978-0-444-63506-8.00004-5.
10.1016/j.cattod.2007.06.082.
[36] G. Engelhardt, Solid state NMR spectroscopy applied to zeolites, Stud. Surf. Sci.
[10] J.C. Groen, W. Zhu, S. Brouwer, S.J. Huynink, F. Kapteijn, J.A. Moulijn, J. Pérez-
Catal. 137 (2001) 387–418, https://doi.org/10.1016/s0167-2991(01)80251-2.
Ramírez, Direct demonstration of enhanced diffusion in mesoporous ZSM-5 zeolite
[37] K.M. Leung, P.P. Edwards, E. Jones, A. Sartbaeva, Microwave synthesis of LTN
obtained via controlled desilication, J. Am. Chem. Soc. 129 (2007) 355–360,
framework zeolite with no organic structure directing agents, RSC Adv. 5 (2015)
https://doi.org/10.1021/ja065737o.
35580–35585, https://doi.org/10.1039/c4ra16583g.
[11] G.V. Shanbhag, M. Choi, J. Kim, R. Ryoo, Mesoporous sodalite: a novel, stable solid
catalyst for base-catalyzed organic transformations, J. Catal. 264 (2009) 88–92,
https://doi.org/10.1016/j.jcat.2009.03.014.

8
J. Floriano da Silva et al. Microporous and Mesoporous Materials 331 (2022) 111640

[38] L. Price, K. Leung, A. Sartbaeva, Local and average structural changes in zeolite A [42] D. Barthomeuf, Framework induced basicity in zeolites, Microporous Mesoporous
upon ion exchange, Magnetochemistry 3 (2017) 42, https://doi.org/10.3390/ Mater. 66 (2003) 1–14, https://doi.org/10.1016/j.micromeso.2003.08.006.
magnetochemistry3040042. [43] A. Corma, R.M. Martín-Aranda, F. Sánchez, Zeolites as base catalysts: condensation
[39] D.W. Breck, Zeolite Molecular Sieves : Structure, Chemistry and Use, Jonh Wiley & of benzaldehyde derivatives with activated methylenic compounds on Germanium-
Sons Inc, Nova York, 1974. substituted faujasite, J. Catal. 126 (1990) 192–198, https://doi.org/10.1016/
[40] M. Smaihi, O. Barida, V. Valtchev, Investigation of the crystallization stages of 0021-9517(90)90057-Q.
LTA-type zeolite by complementary characterization techniques, Eur. J. Inorg. [44] A. Corma, Inorganic solid acids and their use in acid-catalyzed hydrocarbon
Chem. (2003) 4370–4377, https://doi.org/10.1002/ejic.200300154. reactions, Chem. Rev. 95 (1995) 559–614, https://doi.org/10.1021/cr00035a006.
[41] L. Martins, D. Cardoso, Aplicação catalítica de peneiras moleculares básicas micro
e mesoporosas, Quim. Nova 29 (2006) 358–364, https://doi.org/10.2174/
157017806776611962.

You might also like