Vous êtes sur la page 1sur 32

Accepted Manuscript

Alkaline-acid treated zeolite L as catalyst in ethanol dehydration process

Karolina A. Tarach, Justyna Tekla, Urszula Filek, Agnieszka Szymocha, Iwona


Tarach, Kinga Góra-Marek

PII: S1387-1811(16)30594-7
DOI: 10.1016/j.micromeso.2016.12.035
Reference: MICMAT 8064

To appear in: Microporous and Mesoporous Materials

Received Date: 26 October 2016


Revised Date: 15 December 2016
Accepted Date: 31 December 2016

Please cite this article as: K.A. Tarach, J. Tekla, U. Filek, A. Szymocha, I. Tarach, K. Góra-Marek,
Alkaline-acid treated zeolite L as catalyst in ethanol dehydration process, Microporous and Mesoporous
Materials (2017), doi: 10.1016/j.micromeso.2016.12.035.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT
1 Alkaline-acid treated zeolite L as catalyst in ethanol dehydration
2 process
3

4 Karolina A. Taracha,*, Justyna Teklaa, Urszula Filekb, Agnieszka Szymochac, Iwona Tarachd,
5 and Kinga Góra-Mareka*

PT
6

RI
a
7 Faculty of Chemistry, Jagiellonian University in Kraków, Ingardena 3, 30-060 Kraków, Poland

b
8 Jerzy Haber Institute of Catalysis and Surface Chemistry PAS, Niezapominajek 8, Kraków, Poland

SC
c
9 University of Agriculture, Al. Mickiewicza 21, 31-120 Kraków, Poland

U
d
10 Faculty of Chemistry, Nicolaus Copernicus University, Gagarina 7, 87-100 Toruń, Poland
AN
11 Corresponding author:

12 Kinga Góra-Marek kinga.goramarek@gmail.com


M

13 Karolina A. Tarach karolina.tarach@gmail.com


D

14
TE

15 ABSTRACT

16 In the present study, the catalytic dehydration of ethanol to obtain ethylene over efficiently top-
EP

17 down modified zeolites L was investigated at temperature range of 100 to 300 °C. It revealed that the
18 nature of an acidic Si(OH)Al groups determined the catalytic activity of the studied materials. The
C

19 alkaline/acid treated zeolites L accommodating solely the isolated Si(OH)Al groups exhibited the
20 highest catalytic activity and selectivity to ethylene. Modification routes increasing population of the
AC

21 hydrogen bonded Si(OH)Al…O species reduced an activity of the catalysts. The most active catalyst
22 yielded 99% of ethanol conversion having 96% of ethylene yield at 280 oC. Upon the stability test for
23 30 h during the reaction, the modified zeolite L showed neither formation of higher hydrocarbons
24 nor deactivation due to carbon deposition.

25

26 Keywords: zeolite L; ethanol dehydration; acidity; desilication; dealumination; IR spectroscopy

1
ACCEPTED MANUSCRIPT
1 1. INTRODUCTION
2 Ethanol as a one of the most important and commonly used organic substrate that can be easily
3 transformed into numerous compounds such as diethyl ether (DEE), ethylene (C2=) or higher
4 hydrocarbons (C3+) via dehydration process [1, 2]. DEE, the valuable organic solvent and substrate to
5 synthesize of higher ethers [3] is obtained in the exothermic intermolecular dehydration, while the
6 endothermic intramolecular dehydration leads to the formation of ethylene [1]. Ethylene as the
7 significant petrochemical intermediate allows for the further synthesis of numerous products

PT
8 including higher hydrocarbons [4, 5]. In the frame of a future, the dehydration of ethanol seems to
9 have a great importance in the area of renewable sources, due to the fact that ethanol, DEE,

RI
10 ethylene etc. can be manufactured from the alternative feedstock, such as bioethanol [6, 7]. The
11 ethanol dehydration process is effectively catalyzed by various solid acid catalysts: molecular sieves

SC
12 such as zeolites [8, 9] and SAPO materials [10], silica-alumina [9], oxides [11] and heteropolyacids
13 [12]. Zeolites own high catalytic activity in dehydration of ethanol to the specific features: (i)

U
14 micropores system and (ii) presence of strong acid sites, both promote reaction rates and selectivity
15 and alter the stability of surface-bound intermediates [8, 13, 14]. Many efforts were devoted to
AN
16 examining the catalytic behavior of several zeolite structures e.g. H-MOR, H-BEA, H-FER [4, 15], and
17 H-MFI both microporous [16] and hierarchical [2]. Many authors have reported that the fabrication
M

18 of the secondary mesoporosity benefits the coking resistance of hierarchical zeolites [17]. The
19 bottom-up approaches gather the methods dedicated to the creation of mesoporosity at the stage of
D

20 zeolite synthesis thus they require the employment of the specific templates [3, 18-20], proceed
21 through the zeolitization of mesoporous materials [21] or concern direct synthesis of nanosized
TE

22 crystals [17, 22]. Zeolitic systems obtained by these methods accommodate the acid sites of high
23 strength typical of the microporous environment and their enhanced accessibility is provided by well-
EP

24 defined mesoporosity. In top-down approaches, the bimodal porosity is formed by a preferential


25 extraction of the two main components of zeolites, i.e. silica (via desilication) [23-25] or alumina (via
C

26 dealumination) [26]. The steaming of zeolite Y [29, 30] is a well-known example of dealumination
27 process. However, the removal of silicon atoms from zeolite framework has been reported to be one
AC

28 of the most effective approaches for the fabrication of uniform intracrystalline mesoporosity.
29 Desilication process is structure sensitive because the number of Si atoms that could be extracted
30 with the preservation of crystallinity is governed by the features inherent to zeolites: Si/Al ratio and
31 framework topology [27]. Also, the alkaline treatment conditions such as a type and concentration of
32 desilicating agent used are vital factors affecting the acidic property and a uniformity of mesopore
33 system [23, 28]. Generally, the zeolites obtained via desilication route accommodate higher
34 concentration of acidic sites that their solely microporous analogs, however, the strength of protonic
35 moieties is usually reduced. The influence of zeolite topology, hydroxyl group location [8], Lewis and

2
ACCEPTED MANUSCRIPT
1 Brønsted acidity and confinement effects [9, 29] on catalytic ethanol dehydration have been
2 extensively studied. Phung et al. [9] and Chiang and Bhan [8] reported that H-MOR is one of the most
3 effective catalysts that demonstrated selectivity to ethylene in low temperatures and, even, at low
4 ethanol conversion. Among considered catalysts, highly effective, is ZSM-5 zeolite [2], however,
5 microporous characteristic of MFI structure and presence of acid sites of high strength meets the
6 problem of the fast coke precursors formation what ultimately affects activity, selectivity, and
7 lifetime. Also, the problem of bimolecular reactions that lead to the formation of carbon deposit

PT
8 significantly inhibits the catalytic potential of large-pore structures such as H-MOR and H-BEA [9].
9 Due to the fact of high catalytic activity of 12-MR zeolites (H-MOR and H-BEA), we devoted this study

RI
10 to determine the catalytic behavior of other 12-MR zeolite L in ethanol dehydration process.

SC
11 Zeolite L (LTL-framework) is a synthetic, crystalline aluminosilicate synthesized for the first time in
12 the 1950s [30]. The framework Si/Al ratio typically about 3.0 [31] allows classifying L as a low
13 siliceous. Determined by Barrer and Villiger [32] structure of L zeolite consist of a one-dimensional
14
U
network of channels in which access to the internal pore volume is controlled by 12-MR apertures
AN
15 with a diameter of 7.1 Å [33-35]. The unique morphology and structural features of L zeolite affect its
16 performance in numerous applications like separation and catalysis, especially Pt-loaded zeolite L is a
17 distinguished monofunctional catalyst for aromatization of the C6-hydrocarbons [36] commercialized
M

18 in the Chevron Aromax process [15]. Moreover, a number of studies reported that unique pore
19 architecture of L structure perverts the bimolecular reactions and consequently inhibits the
D

20 formation of coke precursors [37]. Thus it is expected that zeolite L can be highly effective, selective
TE

21 (to the ethylene) and stable catalyst for ethanol dehydration process.

22 Till now hierarchical zeolites L were synthesized by various bottom-up strategies, however, materials
EP

23 with limited mesoporosity and poor crystallinity were obtained [38, 39]. The post-synthetic
24 modifications as dealumination [33] and desilication processes were studied in limited extent [40],
C

25 and high reduction in crystallinity and micropore volume was reported [33].
AC

26 In the present study, we attempted to modify the textural and acidic characteristic of zeolite L in
27 order to enhance its catalytic performance in acid catalyzed dehydration process of ethanol. Both
28 features were tuned on the basis of the sequential desilication and dealumination approaches. In
29 desilication, the mesopores are formed by a controlled leaching of the zeolite framework with
30 alkaline solutions. According to literature, the treatment with NaOH solutions is addressed to 10-ring
31 and 12-ring zeolites with 8 < Si/Al < 100 and 4 < Si/Al < 10, respectively [40]. The Al-rich zeolites are
32 however less prone for a direct base treatment due to the protecting role of negatively charged AlO4-
33 tetrahedral [41] and Al distribution in zeolite framework [40]. In this case, the solution is initial

3
ACCEPTED MANUSCRIPT
1 dealumination of the zeolite framework to increase the framework Si/Al ratio and to enable its
2 subsequent dissolution in desilication process. The efficiency of the dealumination has an important
3 impact on the effectiveness of the subsequent. Indeed, the optimized demetallation of the
4 framework leads to the development of extensive mesoporosity and affects the acidic characteristic
5 of zeolite by the alteration of the number and nature of acid sites. The detailed characterization of
6 hierarchical zeolites L was based on ICP OES, XRD, 29Si NMR and low temperature N2 physisorption
7 studies as well as on IR studies of different probe molecules sorption. The results of IR studies

PT
8 allowed to distinguish between different types of acidic hydroxyl groups present in zeolites L and
9 their role in acid catalyzed ethanol dehydration process. The changes of textural and acidic

RI
10 properties find their reflection in catalytic performance of hierarchical zeolites L. Thus the catalytic
11 activity of top-down modified zeolites was investigated in the ethanol dehydration reaction.

SC
12

U
AN
M
D
TE
C EP
AC

4
ACCEPTED MANUSCRIPT
1 2. EXPERIMENTAL SECTION

2 2.1. Catalysts preparation

3 The Na, K-form of zeolite L of Si/Al = 3 purchased from Tosoh (HSZ 500 KOA), hereafter denoted as L,
4 was modified by combining the several routes to obtain the mesoporous analogs (Table S1).

5 The first series (I-s) of materials was obtained by alkaline leaching of zeolite L in 0.8 M NaOH solution

PT
6 at a temperature of 65 and 80 oC for 2 h. After desilication a suspension was cooled down in an ice-
7 bath and filtered. Hierarchically structured zeolites were washed with water until neutral pH. Next,
8 one portion of desilicated materials was leached with 0.2 M HNO3 solution at 65 oC for 2 h (3 g of

RI
9 zeolite per 100 ml HNO3) in order to remove extraframework aluminum species formed during
10 desilication. All modified materials were subjected to threefold ion-exchange with 0.5 M NH4NO3

SC
11 performed at 60 oC for 1 h. Finally, the zeolites were again filtrated, washed and dried at room
12 temperature.

13
U
Other hierarchization procedures, as presented below, were based on the previous removal of Al
AN
14 atoms from the framework of zeolite L via treatment with acidic solutions and calcination in the
15 presence of water vapor. Therefore, the second series (II-s) of hierarchical zeolites L was obtained by
M

16 dealumination of zeolite L with 0.2 M or 0.3 M HNO3 solution at 65 oC for 2 h (3 g of zeolite per 100
17 ml of HNO3) followed by alkaline treatment with 0.1 M NaOH solution at 65 oC for 30 min. (3 g of
D

18 zeolite per 100 ml of NaOH). The third hierarchization procedure (III-s) was based on the
19 dealumination in the mild conditions with 0.2 M and 0.5 M NH4Cl solution at 80 oC for 1 h (3 g of
TE

20 zeolite per 100 ml of NH4Cl). The obtained samples were subsequently desilicated with NaOH
21 solution. Such procedure allowed for a better local pH control during dealumination process and
EP

22 simultaneous of Na+ and K+ for NH4+ ion-exchange. In the last modification route (IV-s) the zeolite L
23 previously ion-exchanged three times with 0.5 M NH4NO3 at 60 oC for 1 h was steamed at a
temperature of 550 oC or 600 oC. Steaming of zeolites L was performed under continuous liquid
C

24
25 water injection with 1.8 cm3min-1 flow speed for 3 hours. The catalysts were preheated from RT to
AC

26 steaming temperature during 2 hours. Subsequently, the desilication with 0.8 M NaOH solution at 80
o
27 C for 2 hours was performed. All modified materials were next subjected to threefold ion-exchange
28 with 0.5 M NH4NO3 performed at 60 oC for 1 h. Finally, the zeolites were filtrated, washed and dried
29 at room temperature. The list of modified zeolites L is given in Table S1 and Figure 1.

30 Additionally, ammonium form of zeolite L, denote as n-L, obtained by threefold ion-exchange


31 procedure with 0.5 M NH4NO3 at 60 oC for 1 h was used as reference material.

32 All materials were calcined at 450 oC for 2 h.

5
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 1. The sequence of post-synthetic treatments over parent n-L to prepare its hierarchical analogs.

U
1 2.2. Materials characterization
AN
2 2.2.1. Chemical analysis, structural and textural properties

3 Si and Al concentrations in the parent and desilicated zeolites were determined by the ICP OES
M

4 method with an Optima 2100DV (PerkinElmer) spectrometer.

5 Powder X-ray diffraction (XRD) measurements were carried out using a PANalytical Cubix
D

6 diffractometer, with CuKα radiation, λ=1.5418 Å and a graphite monochromator in the 2θ angle range
TE

7 of 5-40°. X-ray powder patterns were used for structural identification of the relative crystallinity
8 values (%Cryst) for all the zeolites. The determination of the relative crystallinity value was based on
9 the intensity of the characteristic peaks in the range between 22° and 33°.
EP

10 N2 sorption processes at -196 oC were studied on an ASAP 2420 Micromeritics after activation in
C

11 vacuum at 400 oC for 12 h. Surface area (SBET) and micropore volume (Vmicro) were determined by
12 applying the BET and t-plot methods, respectively. The volume of mesopores (Vmeso) was obtained by
AC

13 applying the BJH model to the adsorption branch of the isotherm. The mesopore surface area (Smeso)
14 was calculated in the range between 2 and 30 nm with BJH model.

15 2.2.2. MAS NMR studies

16 Solid state MAS NMR spectra were acquired on an APOLLO console (Tecmag) at the magnetic field of
17 7.05 T (Magnex). In the case of the 29Si MAS-NMR spectra a 3 μsrf pulse (π/2 flipping angle) was
18 applied, 4 kHz spinning speed, and 256 scans with the delay of 20 s were acquired. Chemical shifts

6
ACCEPTED MANUSCRIPT
1 are reported in ppm relative to an external standard of TMS for 29Si. The MAS NMR spectra were
2 normalized to the same mass of the sample.

3 2.2.3. IR studies

4 Prior to the FTIR study, the self-supporting wafers (ca. 5-10 mg·cm-2) of the zeolitic materials were
5 pre-treated in situ in homemade quartz IR cell at 450 oC under vacuum conditions for 1 hour. The IR
6 spectra were recorded with a Vertex 70 spectrometer equipped with an MCT detector. The spectral

PT
7 resolution was 2 cm−1.

RI
8 The concentration of Brønsted and Lewis acid sites was obtained in quantitative IR studies of pyridine
9 (Avantor Performance Materials Poland S.A.) adsorption according to the procedure previously
10 reported [42-44]. The sample was saturated with pyridine (10 Tr in the gas phase) at 170 oC and next

SC
11 evacuated at the same temperature to remove the gaseous and physisorbed Py molecules. Then, the
12 IR spectrum was taken at a temperature of 170 oC. The concentration of Brønsted and Lewis sites

U
13 was calculated using respectively the intensities (the heights) of the 1545 cm−1 band of the PyH+ ions
AN
14 and the 1450 cm−1 PyL band of Py coordinatively bonded to Lewis sites as well as the values of the
15 extinction coefficient of 0.07 cm2 μmol-1 for the PyH+ band and 0.10 cm2 μmol-1 for the PyL.
M

16 The acid strength was determined based on Py thermodesorption and carbon monoxide adsorption
17 studies. In the Py-thermodesorption experiments, the conservation of the 1545 cm−1 band of the
D

18 PyH+ ions and the 1450 cm−1 PyL bands under the desorption procedure at 330 oC were taken as a
19 measure of the acid strength of the Brønsted and Lewis sites, respectively. The sorption of CO (Linde
TE

20 Gas Poland, 99.95 %) as probe molecule was performed at -130 oC. The downshift of IR band of the
21 acidic Si(OH)Al hydroxyls due to their hydrogen bonding to CO molecules was taken as a measure of
EP

22 the acid strength.

23 The sorption of ethanol (Sigma-Aldrich, absolute, ≥ 99.8%) was performed at room temperature, and
C

24 next the temperature was increased to 150 oC. Then the sample was cooled down to room
AC

25 temperature and the spectrum was collected.

26 2.2.4. Raman Spectroscopy

27 The microRaman analysis was performed with a Renishaw InVia dispersive spectrometer equipped
28 with a CCD detector and integrated with a Leica DMLM confocal microscope. An excitation
29 wavelength of 514.5 nm was provided by an Ar-ion laser (Spectra-Physics, model 2025). The spectra
30 were recorded at ambient conditions with the resolution of 2 cm−1. The Raman scattered light was
31 collected with a 50x Olympus objective in the spectral range of 1000–2000 cm−1.

7
ACCEPTED MANUSCRIPT
1 2.3. Ethanol dehydration reaction

2 Activities of the investigated samples were measured in ethanol dehydration in the vapor phase. For
3 each experiment, 0.05 g of fresh catalyst was mixed with 0.5 g of quartz grains and placed in a
4 differential microreactor connected online with a gas chromatograph. Reaction products were
5 analyzed using a PerkinElmer Clarus 580 equipped with Elite-Plot Q capillary column (30 m of length,
6 0.53 mm of inner diameter) and TCD detector. Prior the catalytic reaction the samples were

PT
7 standardized under continuous helium flow for 30 min at 300 oC. Then helium as a carrier gas was
8 passed over the liquid ethanol obtaining the concentration of ethanol in helium flow 0.56 mmol·dm-3.
The total flow rate of the feed stream was kept at 35 cm3·min-1. The reaction temperature was

RI
9
10 changed from 100 to 300 oC. The space velocity in performed catalytic test was kept at 1.7
gethanol/(gcatalyst·h). The calculation of TOF (turnover frequency / s-1) values was based on equation 1

SC
11
12 given in ref [45].

U
Reactant flowrate μmol s-1 ∙Conversion
TOF s -1 =
Quantity of sites μmol g -1 !∙Catalyst weight g!
AN
13 eq. (1)
M
D
TE
C EP
AC

8
ACCEPTED MANUSCRIPT
1 3. RESULTS AND DISCUSSION

2 3.1. Chemical analysis, structural and textural properties

3 The data gathered in Table 1 provide information on the Si/Al ratios of the conventional zeolite L and
4 its modified analogs in accordance with previously listed modification procedures. The ion-exchange
5 procedure seems to lead to a minor dealumination as the Si/Al ratio slightly increased (n-L sample).
6 In the case of samples modified via I route a slight decrease in the value of the Si/Al ratio can be

PT
7 noticed for both materials subjected to desilication using NaOH solution (samples L_I-a and L_I-b). The
8 number of Si atoms extracted from the zeolitic framework was not enhanced by the increase of the

RI
9 temperature pointing to the low efficiency of the alkaline leaching. This effect is associated with the
10 presence of a large number of aluminum atoms in the structure of zeolite L. The Al-rich

SC
11 aluminosilicates are less prone to desilication since the negative charge of AlO4- tetrahedra repel
12 charged OH- ions, thus the Si-O-Al bond is poorly susceptible to hydrolysis in the presence of hydroxyl
13 ions. As a consequence, the presence of framework Al atoms prevents the neighboring silicon atoms
14
U
from the extraction and the desilication process is effectively inhibited [46]. The treatment of
AN
15 desilicated zeolites with nitric acid(V) (HNO3) resulted in the expected increase in Si/Al ratio values
16 (sample L_I-c and L_I-d) due to (i) the partial dealumination of the zeolite framework and (ii) the
M

17 removal of aluminium extraframework material, AlEF, previously accumulated in the zeolitic channels
18 or on mesopore surface upon desilication process. The XRD patterns (Fig. 2A) of the materials
D

19 obtained via I route are typical of zeolite L [47, 48]. The demetallation processes did not affect the
20 crystallinity of the modified materials (Table 1), no presence of amorphous phase was confirmed.
TE

21 Noticeable, however, small differences in the intensity of the corresponding reflections can be
22 associated with the changes in the size of the elemental cells as the expected consequence of
EP

23 demetallation and ion-exchange processes.

24 The opposite sequence of desilication and dealumination processes (II route) resulted in the
C

25 expected substantial alternation in the values of the Si/Al ratio. The samples subjected to nitric
26 acid(V) (HNO3) treatment were characterized by significantly increased Si/Al ratios (Si/Al = 6.0 for
AC

27 L_II-a and Si/Al = 8.5 for L_II-b). Subsequent alkaline leaching resulted in the enrichment of zeolite
28 crystals in Al-atoms; the Si/Al ratios decreased to 4.3 and 8.3 for the L_II-c and L_II-d, respectively.
29 Such observation leads to the conclusion that the removal of aluminum atoms from zeolite
30 framework significantly improves the efficiency of the desilication process. Nevertheless, as can be
31 stated from the XRD patterns (Fig. 2B) the dealumination of zeolite n-L in more severe conditions (0.3
32 M HNO3, L_II-b) caused its partial amorphization. Zeolites treated in milder conditions (0.2 M HNO3,
33 L_II-a) retained the structure of the native zeolite L and were characterized by high Si/Al ratio.

9
ACCEPTED MANUSCRIPT
1 Subsequent desilication process allowed for retaining the crystallinity and micropore volume of
2 modified zeolites.

3 The third modification path related to the treatment of zeolite n-L with NH4Cl solution and, in
4 consecutive step, the desilication with NaOH offered only slight changes in aluminum content in the
5 modified zeolites (Table 1). The XRD patterns (Fig. 2C) point to the preservation of full crystallinity of
6 zeolitic materials (L_III-a – d). The samples previously calcined in the presence of water vapor and

PT
7 subsequently desilicated with NaOH (IV route) gave rise to Si/Al ratio (Table 1). It can suggest that for
8 these materials both framework silicon atoms and aluminum extraframework species are prone to

RI
9 the alkaline leaching to the same extent. Such phenomenon can be favored in highly disturbed by
10 steaming process zeolite structure. The analysis of the diffractograms of the materials modified by

SC
11 steaming IV route revealed significant aberrations in the pattern of the zeolite steamed at 600 oC due
12 to the zeolite structure destruction (Fig. 2D). Lowering of the temperature of steaming process to
13 550 oC allowed for the retaining crystallinity of zeolite L. After desilication the reflections are shifted
14
U
to higher values of 2θ pointing to the increase of unit cell parameters due to the removal of Si atoms
AN
15 from the framework.

A L_I_d
B L_II_d
M

L_I-c L_II-c
D

L_I-b L_II-b
TE

L_I-a L_II-a

n-L n-L
EP

10 15 20 25 30 35 40 10 15 20 25 30 35 40
2Θ / ο 2Θ / ο
C L_III_d D
C

L_IV-c
L_III-c
AC

L_IV-b
L_III-b

L_IV-a
L_III-a

n-L n-L

10 15 20 25 30 35 40 10 15 20 25 30 35 40
2Θ / ο 2Θ / ο
Fig. 2. The XRD patterns of zeolites L modified according to procedures I (A), II (B), III (C), and IV (D).

16

10
ACCEPTED MANUSCRIPT
1 Table 1. Relative crystallinity values (%Cryst) derived from XRD, the composition determined by both chemical analysis
29
2 (Si/AlICP) and Si MAS NMR (Si/AlNMR), as well as the textural parameters from low-temperature N2 adsorption/desorption
3 studies for the native zeolite L and hierarchical ones.

%Cryst Si/Albulk Si/AlNMR SBET Smikro Smeso Vmicro Vmeso


Sample 2 2 2 3 3
% a.u. a.u. m /g m /g m /g cm /g cm /g
L 100 2.9 2.7 405 347 58 0.14 0.08
n-L 90 3.0 2.7 433 367 66 0.15 0.09
L_I-a 89 2.8 2.5 424 388 36 0.15 0.05

PT
L_I-b 98 2.7 2.6 426 379 47 0.15 0.07
L_I-c 79 3.7 3.8 455 365 90 0.15 0.11
L_I-d 83 3.7 3.3 420 342 78 0.15 0.11

RI
L_II-a 59 6.0 5.7 483 415 69 0.17 0.10
L_II-b 46 8.5 7.0 445 413 32 0.16 0.05

SC
L_II-c 92 4.3 4.1 511 377 134 0.16 0.21
L_II-d 56 8.3 7.9 391 256 135 0.11 0.19
L_III-a 99 3.0 2.9 451 419 32 0.16 0.06
100 3.1 3.1 448 423 25 0.16 0.04

U
L_III-b
L_III-c 100 3.3 3.0 450 421 29 0.16 0.05
AN
L_III-d 100 3.2 3.1 465 407 59 0.16 0.09
L_IV-a 19 2.7 nd 115 91 25 0.04 0.04
L_IV-b 32 2.8 2.8 202 170 32 0.07 0.09
M

L_IV-c 46 3.4 4.4 317 269 48 0.11 0.07

4 Textural properties of studied zeolites are summarized in Table 1. N2 adsorption/desorption isotherm


D

5 of the parent zeolite n-L corresponds to a type-I in the IUPAC classification. Together with
6 mesoporosity enhancement for the modified zeolites L an isotherm type-IV of micro/mesoporous
TE

7 materials is developed (Figure S1), especially well seen for materials treated according to second
8 route (Figure S1 B). The conventional zeolite L showed a solely microporous structure (Vmicro = 0.14
EP

9 cm3·g-1 ) with only small mesoporous surface resulting from interparticle voids. The slightly increased
10 micropore volume after the ion-exchange procedure should be related to the removal of potassium
C

11 cations from zeolite pores. The sequentially modified samples exhibited a hierarchical structure of
12 different extent of mesoporous surface depending on the type of applied treatment. The most
AC

13 effective for mesopores generation was second route involving sequential dealumination/desilication
14 treatment in relatively severe conditions (0.2 M HNO3). Application of mild conditions of
15 dealumination (0.2 M NH4Cl) in third modification route limited the mesopores creation upon
16 subsequent desilication and only minor Smeso values were found. The micropore volume of modified
17 samples followed the trend of changes observed for values of crystallinity. The preservation of
18 structure of modified zeolites in I – III routes allowed for maintenance of micropore volumes on a
19 similar level. The drop of the micropore volume for samples modified in the fourth route should be
20 assigned either to deposition of extra-framework material inside the micropores or to the loss of

11
ACCEPTED MANUSCRIPT
1 crystallinity, as mentioned above. Properly chosen sequential dealumination/desilication procedure
2 enhanced the mesopore surface area to 134 m2·g-1 and preserved micropore volume and
3 crystallinity.

4 3.2. 29Si MAS NMR studies

5 The course of demetallation was monitored by 29Si MAS NMR and quantitatively estimated as the
6 alteration of the Si/Al framework ratio (Si/AlNMR, Table 1). The 29Si MAS NMR spectra (normalized to

PT
7 the sample mass, Figure 3) of native zeolite n-L and modified zeolites are a superposition of several
8 signals. After deconvolution, the four lines at −92, −97, −102 and -107 ppm were assigned to

RI
9 Si(1Si, 3Al), Si(2Si, 2Al), Si(3Si,1Al) and Si(4Si,0Al) sites, respectively. Interestingly, desilication of
10 zeolite n-L according I route resulted in the appearance of the -92 ppm signal of Si(4Si,0Al) sites,

SC
11 however, preferential removal of any type of aluminum atoms was not observed. It has been
12 reported that dealumination and desilication processes, both resulting in mesopore formation,

U
13 requires an internal migration of T atoms (T= Si, Al) filling up the framework vacancies (T-jump
14 mechanism) [49, 50]. The appearance of the Si(0Si,4Al) sites upon desilication was assigned to such
AN
15 healing process in which lattice vacancies created during the desilication were filled up with Al
16 atoms. While the Al dissolved species prefer to be readsorbed onto zeolite surfaces the dissolved Si
M

17 species aggregate in solution, and above mentioned process is fully justified [51]. In subsequent
18 dealumination procedure, the Al atoms are extracted from those newly formed Si(0Si,4Al) sites to
D

19 produce Al-free zones detected as the Si(4Si,0Al) sites (-107 ppm line) and (SiO)4Si defects/an
20 amorphous component (-112 ppm line).
TE

21 During severe dealumination procedures (II and IV) the Si atoms were preferentially extracted from
22 Si(1Si, 3Al) and Si(2Si, 2Al) sites and the formation of an amorphous component, which is
EP

23 represented by a broad line at –112 ppm was observed. Subsequent desilication did not perturb the
24 intensity of the Si(2Si, 2Al) and Si(3Si,1Al) signals confirming a higher stability of Si atoms placed in
C

25 the neighborhood of the aluminium enriched units over Si atoms in Si(4Si,0Al) units. Similar results
AC

26 were reported in our earlier study on the desilication of zeolite ZSM-5 [44, 52] where it was shown
27 that the AlO4- stabilizes neighboring Si atom by the repulsion of OH− ions. When acid leaching was
29
28 carried out in mild conditions (III-route) no significant changes in Si MAS NMR spectra were
29 detected.

30

31

12
ACCEPTED MANUSCRIPT

29 29 29 29
Si NMR Si NMR Si NMR Si NMR
4 kHz MAS A 4 kHz MAS B 4 kHz MAS C 4 kHz MAS D
absolute scale absolute scale absolute scale absolute scale
Si(3Si,1Al) Si(3Si,1Al)
Si(3Si,1Al)

PT
Si(2Si,2Al) Si(2Si,2Al)
Si(4Si,0Al)
(SiO)4Si Si(1Si,3Al) Si(3Si,1Al)
Si(4Si,0Al) Si(2Si,2Al)
Si(1Si,3Al)
Si(4Si,0Al) Si(2Si,2Al)
(SiO)4Si (SiO)4Si
Si(0Si,4Al) Si(1Si,3Al)
L_I-d L_II-d L_III-d Si(3Si,1Al)

RI
Si(1Si,3Al)

SC
L_I-c L_II-c L_III-c L_IV-c

U
L_I-b L_II-b L_III-b

AN
Si(4Si,0Al) (SiO)4Si
L_IV-b

M
L_II-a L_III-a
L_I-a

D
n-L n-L n-L n-L

TE
-60 -80 -100 -120 -140 -60 -80 -100 -120 -140 -60 -80 -100 -120 -140 -60 -80 -100 -120 -140
ppm from TMS ppm from TMS ppm from TMS ppm from TMS
EP
29
Fig. 3. The Si MAS NMR spectra of zeolites L modified according to procedures I (A), II (B), III (C), and IV (D). Deconvoluted spectra are presented for materials modified by I-route (A).
C
AC

13
ACCEPTED MANUSCRIPT
1 3.3. Acidic properties

2 3.3.1. The Si(OH)Al groups in zeolites L

3 The spectra of the hydroxyls groups in n-L and modified zeolites L are presented in Fig. 4A. The
4 3710 cm-1 and 3740 cm-1 bands were assigned to the Si-OH in defects and Si-OH freely vibrating on
5 external surfaces and mesopore surfaces, respectively. Apart from non- or weakly acidic silanols
6 groups also the bridging Si(OH)Al groups are populated in micropores of zeolites L. The presence of

PT
7 3680 cm-1 bands assigned to extraframework non-acidic aluminum species points to low thermal
8 stability of aluminum atoms in the tetrahedral position of zeolites L. The 3630-3640 cm-1 band is

RI
9 typical of the Si(OH)Al groups freely vibrating in 12-MR channels (hereafter denoted as type A, Fig.
10 5A) while a broad band with a maximum centred at 3250 cm-1 originates from the bridging hydroxyl

SC
11 groups in smaller cavities influenced by an additional interaction with the zeolitic framework, i.e.
12 Si(OH)Al…O species (hereafter denoted as type B, Fig. 5B). The existence of such strong hydrogen
13 bonded Si(OH)Al…O species has been reported for the first time by Zholobenko et al. [53] in zeolites
14
U
ZSM-5 and mordenites, and further confirmed by 1H MAS NMR studies [54].
AN
15 Among all zeolites studied, the relatively high population of strong hydrogen bonded Si(OH)Al…O
16 species (the type B) is typical of zeolite n-L. It is manifested by the highest intensity of the 3250 cm-1
M

17 band as well as the ABC structure being a characteristic feature in the spectra of very strongly
18 hydrogen-bonded systems (Fig. 4A). The corresponding minima between the three submaxima are
D

19 situated at 2557 and 1940 cm-1. Demetalation procedures strongly influenced the relative ratio
TE

20 between isolated Si(OH)Al (the type A) and hydrogen bonded hydroxyls Si(OH)Al…O (the type B). The
21 latter species are effectively eliminated by strong acid treatment with HNO3 (II route) or consecutive
22 alkaline and acid treatment (the route I – samples L_I-c and -d). However, during mild acid and
EP

23 alkaline leaching, the population of these species either is significantly enhanced or remains at the
24 same level. The different origin of these hydroxyls gives the impact to the acidic characteristics and
C

25 finally to the catalytic performance. Therefore the ability of the hydroxyls to be involved into
AC

26 interaction with the probe molecules differing in basicity was examined. The zeolite n-L was chosen
27 as the reference due to high population both the type A and the type B species. Sorption of pyridine
28 on n-L resulted in the consumption of hydroxyls of both types (Fig. 6A), being manifested by the
29 negative bands at the frequencies of 3635 cm-1 and 3300 cm-1. It evidences that adsorption of Py
30 molecules obviously resulted in the cancellation of the interaction between the part of the hydroxyl
31 groups and the zeolitic framework. Also, ABC hydrogen bond structure was partially destroyed.

32
33

14
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 4. (A) The IR spectra of studied zeolites in the region of O-H groups vibration. (B) The ABC structure in the IR spectrum
of zeolite n-L of the hydrogen bonded Si(OH)Al…O groups.
1

Fig. 5. The models of isolated (A) and hydrogen bonded (B) Si(OH)Al hydroxyls.
2
3

15
ACCEPTED MANUSCRIPT
A 3630 3250
Py A=0.1
3745 PyH+
a 1545 PyL
1450
b

c=b-a c=b-a

e=d-a e=d-a
3600 3200 2800 2400 1600 1500 1400

PT
ν [cm ]
-1

B CO A=0.1

RI
3630 3250
3745 2557

SC
b
∆ν OH....CO
c=b-a
3800 3600 3400 3200 3000 2800 2600 2400

U
ν [cm-1]
AN
C Si(OH)Al Si(OH)Al...O Et-OH A=0.1
Si-OH 3630 3250
2557
3745 a
b
M

c = b -a

2507 d
D

e=d-a
TE

3800 3600 3400 3200 3000 2800 2600 2400


ν [cm-1]
Fig. 6. The IR spectra of the hydroxyl groups in zeolite n-L perturbed by the interaction with:
EP

o
(A) pyridine: a – activated zeolite, b – after saturation of all accessible hydroxyls at 170 C, c – the difference spectrum (b-a),
o o
d – after desorption at 330 C and cooling down to 170 C, e – the difference spectrum (d-a).
o
(B) carbon monoxide: a – activated zeolite, b – after saturation all accessible hydroxyls at -130 C, c – the difference
C

spectrum (b-a).
(C) – ethanol: a – activated zeolite, b – after saturation of all accessible hydroxyls at room temperature, c – the difference
AC

o
spectrum (b-a), d – after saturation of all accessible hydroxyls at 150 C and cooling down to RT, e – the difference spectrum
(d-a).

1 Desorption of Py at 330 oC leading to the partial decomposition of PyH+ ions (40%) resulted in the
2 restoring only the band of the type B groups, the 3635 cm-1 band of type A hydroxyl was not rebuilt
3 (Fig. 6A). The opposite situation was observed for the low-temperature sorption of CO on n-L (Fig.
4 6B). Weakly basic CO molecules were able to interact solely with isolated Si(OH)Al groups (the type
5 A) while hydroxyls engaged into interaction with the zeolitic framework, i.e. the type B remained
6 unperturbed. No changes were observed in ABC hydrogen bond structure (the bands at 3250 and

16
ACCEPTED MANUSCRIPT
1 2557 cm-1) upon the sorption of an excess of carbon monoxide. Sorption of the first doses of ethanol
2 in zeolite n-L at room temperature (Fig. 6C) resulted in the consumption of the 3630 cm-1 of the
3 isolated Si(OH)Al groups band in the first order. The 3250 cm-1 band of the type B hydroxyls was only
4 negligibly perturbed upon admittance of the excess of ethanol. When temperature increased to 150
o
5 C the vanishing of the 3250 cm-1 was more evident, however, only partially consumed. The band
6 typical of ABC hydrogen bond structure were only slightly reduced in their intensities but significantly
7 shifted to lower frequencies. Assuming that all acidic hydroxyls are accessible for ethanol molecules

PT
8 it can be concluded that ethanol is not sufficiently strong base to cancel the hydrogen bond in the
9 type B hydroxyls at a temperature lower than 150 oC. Thus finally the isolated species only are able to

RI
10 bond ethanol and form ethoxy species. It cannot be excluded that the hydroxyls of type B can be
11 involved into interaction with ethanol at higher temperatures. Usually, in the IR spectra of

SC
12 hierarchical zeolites the development of mesopore surface is detected as the increase of the
13 intensity of the silanols band, both isolated (3740 cm-1) and those in defects (3710 cm-1). Such

U
14 behavior was observed only for zeolites modified in I-s and II-s routes. Firstly, the desilication process
15 of zeolite L always leads to increasing the amount of hydrogen bonded Si(OH)Al hydroxyls. The only
AN
16 accessible Si(OH)Al groups (isolated or hydrogen bonded) are located on the wall of the main 12-MR
17 channel, more specifically at the front of a boat shaped eight-membered ring. Furthermore, both, 29Si
M

18 NMR [55] and IR spectroscopy work [56] evidenced an ordered distribution of Al ions, which implied
19 a double unit cell having two different types of sites on the wall of the main channel differing in the
D

20 number of neighboring Al framework atoms equal to one or three (and corresponding negatively
21 charged oxygen atoms) [Delgado, 2015, Infrared spectroscopic and thermodynamic assessment of
TE

22 extraframework cationic adsorption sites in the zeolite K-L by using CO as probe molecule]. Thus, it
23 can be anticipated that observed hydrogen bonded Si(OH)Al…O species (the type B) are mainly
EP

24 present at the front of a boat shaped eight-membered ring with a higher number (three) of Al
25 framework atoms. This environment favors the formation of hydrogen bond between the OH group
C

26 and the framework oxygen or between the neighboring Si(OH)Al groups also situated in eight-
27 membered ring similarly as in the case of silanol nests. The extent of ABC structure being a
AC

28 characteristic feature in the spectra of very strongly hydrogen-bonded systems points to a significant
29 population of Si(OH)Al…O species. During desilication, the hydrogen bonded Si(OH)Al…O species are
30 preserved since the negative charge of three neighboring AlO4- tetrahedra prevents the framework
31 against dissolution with charged OH- ions. In overall, a removal of eight-membered rings
32 accommodating only one framework Al ion is favored thus the amount of the type B species is
33 increased. These hydrogen bonded Si(OH)Al…O species are stable enough to not be perturbed by the
34 weakly basic reagent molecules; ethanol molecules are not able to discard the hydrogen bond even
35 at a temperature as high as 150 oC (Fig. 6). Therefore it is believed that the increasing population of

17
ACCEPTED MANUSCRIPT
1 the type B species modify the sorption property, and further, the catalytic properties of desilicated
2 zeolites L.
3 Contrary to other zeolitic materials, there are a few reports on tuning zeolite L by post-synthetic
4 modifications. Verboekend et al. [40] reported, that a conventional direct desilication did not provide
5 any modification of zeolite L, only a tandem acid-base treatment enabled to obtain hierarchical
6 analogs. However, in this work, we demonstrated that the base treatment affected the nature of the
7 hydroxyls and led to an increase in the population of hydrogen bonded hydroxyls Si(OH)Al…O.

PT
8 The results allow for concluding that dealumination process leads to (i) depletion of hydrogen
9 bonding between Si(OH)Al hydroxyl either framework oxygen or neighboring Si(OH)Al group also

RI
10 situated in eight-membered ring (reduction or vanishing of 3250 cm-1 band seen for all dealuminated
11 samples) and (ii) partial removal of the isolated Si(OH)Al hydroxyls. Dealumination followed by

SC
12 desilication (the route I) as well as a mild dealumination treatment (route III) offered the resulting
13 materials of higher activity than materials alkali-leached only or firstly acid- and then alkali-treated.

U
14 Only strong dealumination of parent zeolite Na,K-L with HNO3 (route II, samples L_II-a and L_II -b) led
15 to decrease in activity, which was again not restored upon desilication. Thus, the increasing
AN
16 population of the hydrogen bonded Si(OH)Al…O species is supposed to have an impact on the
17 catalytic properties of modified zeolites L. The relationship between the number of the eight-
M

18 membered rings in the zeolitic structure, as those delivering the confinement effect and the catalytic
19 activity was widely reported [57]. It has been shown that the steric constraints in the eight-
D

20 membered rings of H-FER significantly decreased a stability of ether (DEE), leading to its fast
21 decomposition thus higher alkene selectivity [ref]. Also, the preferred formation of ethene in H-MOR
TE

22 was attributed to the presence of small eight-membered ring side pockets which prevented the
23 formation of bulky dimer species, favoring monomolecular dehydration and producing ethene [58].
EP

24 Apparently, Al extraction from Na,K-L during the strong dealumination with HNO3 can be responsible
25 for the destruction of eight-membered rings which might be the key to lowered catalytic activity. This
C

26 assumption can be further justified by vital alterations in the chemical composition of strongly
27 dealuminated zeolites (route II), the values of Si/Al ratio increased from 3.0 to 6.0 or 8.5 for 0.2 M
AC

28 and 0.3 M HNO3-leached samples, resp., what corresponds to the loss of 40-60% Al atoms from
29 zeolite framework.

30 3.3.2. Acidic characteristic

31 Pyridine (Py) molecule is widely used as a probe for quantification of both Brønsted and Lewis sites in
32 solid catalysts [42, 43]. The pyridine ring vibration 19b mode is mostly affected by the nature of
33 intermolecular interaction via the nitrogen lone pair electrons. The band ca. 1445 cm-1 originates
34 from pyridine coordinatively bonded to Lewis sites (PyL) while the vibrations of pyridinium ions PyH+

18
ACCEPTED MANUSCRIPT
1 appear as 1545 cm-1 band. It is generally accepted that there is no significant change in the values of
2 the extinction coefficients determined for the bands of Py chemisorbed with respect to the zeolitic
3 structure. The concentrations of both Brønsted (PyH+) and Lewis (PyL) acid sites in studied zeolites
4 (Table 2) were calculated on the basis of the maximum intensities of the PyH+ and PyL bands and the
5 corresponding extinction coefficient values [42, 43]. The information on the acid strength of
6 Brønsted acid sites was obtained in the Py-thermodesorption experiments (the values A330/A170) and

PT
7 low-temperature CO sorption studies (∆νOH…CO).

8 Transformation of Na,K-L into NH4-form (zeolite n-L) involving also mild dealumination led to the

RI
9 generation of Brønsted acid sites in the amount of 31% of aluminum content derived from ICP
10 analysis (Table 2). These acid sites are possibly generated by the aluminum atoms located on the wall

SC
11 of the main 12-MR channel; the crystalline structure of zeolite was retained. The amount of acidic
12 group corresponds to the potassium ions that can be easily removed without distorting the zeolite
13 framework. Thus it concerns ions located on the walls of the 12-MR pore channels or D positions. In
14
U
the I route alkaline leaching with NaOH increased the protonic site's concentrations. Subsequent
AN
15 dealumination reduced the amount of Brønsted acid sites to the half of those present in the n-L
16 sample. Independently from the type of treatment, the global acid strength increased with Si/Al
M

17 ratio. In the II-s and III-s routes differing in the extension of dealumination process the obvious
18 reduction of protonic sites concentration was observed (compare –a and –b denoted samples).
19 Desilication carried out in the next order allowed to rebuilt the protonic acidity. The most significant
D

20 changes were observed for the L_III-d material. Again, subsequent demetallation processes lead to
TE

21 somewhat higher strength of protonic sites with respect to zeolite n-L. Steamed samples (L_IV
22 samples), in spite of the partially preserved crystallinity, demonstrated the drastic decrease of
EP

23 protonic sites concentration followed by a significant enhancement in their acid strength.


24 Interestingly, the samples revealing the highest concentration of protonic sites can be described also
25 as those hosting the highest population of the Si(OH)Al…O species (band 3250 cm-1). Furthermore,
C

26 both mild acid treatment, as well as desilication process, lead to the development of this type of
AC

27 acidity. Demetalation process leading to the deterioration of the zeolite framework is believed to
28 make the ion-exchange of the cations located between cancrinite cavities more facile, thus it offered
29 the formation of the acidic hydroxyls engaged into interaction with framework oxygen atoms
30 Si(OH)Al…O species (the type B). Still, however, the quantitative IR studies of pyridine sorption
31 reflected the limited accessibility of acidic sites in all samples allowing to detect only 25 – 60 % of the
32 possible amount of sites based on aluminum content.

-1 -1
33 Table 2. Content (in µmol·g and on unit cell) of Al atoms from chemical analysis AlICP, the concentrations (in µmol·g ) of
34 Brønsted (B) and Lewis acid sites (L) from IR spectroscopy measurements with pyridine in the parent zeolite L and

19
ACCEPTED MANUSCRIPT
1 hierarchical ones as well as the parameters characterising the acid strength of the Brønsted acid sites achieved from the IR
2 studies of pyridine thermodesorption (expressed as A330/A170) and a low temperature CO sorption (expressed by ∆νO-H···CO in
−1
3 cm ).

+
AlICP AlICP PyH PyL CB+CL A330/A170 ∆vOH…CO
Sample -1 -1 -1 -1 -1
µmol·g u.c. µmol·g µmol·g µmol·g - cm
n-L 3147 8.92 964 360 1324 0.60 310
L_I-a 3350 9.50 1206 266 1472 0.70 333

PT
L_I-b 3447 9.77 1112 194 1306 0.63 327
L_I-c 2703 7.66 410 156 566 0.66 297
L_I-d 2704 7.66 395 166 561 0.68 295

RI
L_II-a 1807 5.12 568 280 848 0.95 318
L_II-b 1343 3.81 297 225 522 0.68 n.d.
L_II-c 2396 6.79 657 477 1134 0.61 312

SC
L_II-d 1364 3.87 455 176 681 0.70 306
L_III-a 3151 8.93 930 173 1103 0.69 312
L_III-b 3121 8.85 875 355 1230 0.68 313

U
L_III-c 2924 8.29 1018 208 1226 0.65 325
L_III-d 3052 8.65 1404 298 1702 0.70 327
AN
L_IV-a 3402 9.64 158 95 253 1.00 340*
L_IV-b 3481 9.87 420 256 676 0.98 325
L_IV-c 2895 8.21 445 204 649 1.00 335
M

4 * the amorphized sample

5 3.3.3. Catalytic behavior in ethanol dehydration


D

6 The catalytic activity of the investigated zeolites in dehydration of ethanol in the vapor-phase was
TE

7 studied at the temperatures ranging from 100 oC to 300 oC under atmospheric pressure, with space
8 velocity equal to 1.7 h-1. Before the catalytic test, the zeolites were in situ heated at 350 oC for 30 min
EP

9 to remove water from zeolite channels. The catalytic performance of the modified materials was
10 compared with the bulk zeolite with regard to the type of generated mesoporosity and retained
11 acidity.
C
AC

12 The zeolite n-L represents the moderate activity in ethanol dehydration reaction, the T50%
13 (temperature needed for 50 % conversion) is ca. 250 oC. The comparison of the conversion curves
14 reveals that the order of the alkali and acid treatment strongly affected the catalytic performance of
15 the modified materials (Fig. 7). The samples dealuminated in the first stage of modification i.e.
16 dealuminated with acid solutions (L_II-a and L_II-b) and steamed at 550 oC (L_IV-b sample) were
17 found to offer the activity comparable to zeolite n-L or were less active. The highly developed surface
18 together with the high concentration of protonic sites did not assure better catalytic performance. It
19 unequivocally evidences that the transformation of small molecules ruled by the confinement factor

20
ACCEPTED MANUSCRIPT
1 is strongly reduced. The loss of 40-60% Al atoms from zeolite framework observed for the strongly
2 dealuminated zeolites (route II) must be reflected in vital structure alternation. Therefore, it is
3 supposed that the destruction of Al-rich eight-membered rings is revealed as lowered catalytic
4 activity of HNO3-treated samples (route II). Subsequent desilication did not provide any significant
5 changes in the activity of the respective materials (L_II-c and -d; L_III-d; L_IV-c). Only the materials
6 treated with NaOH and then leached with HNO3 (I-s route) assured the important activity
7 enhancement on whole temperature range; the T50% was lowered to 200-220 oC. The highest activity

PT
8 was observed for the samples L_I-c and L_I-d.

RI
100 A 100 B
80 80
Conversion /%

Conversion /%

SC
60 60
n-L n-L
40 L_I-a 40 L_II-a
L_I-b

U
20 L_I-c 20 L_II-c
L_I-d L_II-d
0 0
AN
140 160 180 200 220 240 260 280 300 140 160 180 200 220 240 260 280 300
Temperature /oC Temperature /oC
100 C 100 D
M

80 80
Conversion /%

Conversion /%

60 60
D

40 n-L 40
n-L
TE

20 L_III-d 20 L_IV-b
L_III-b L_IV-c
0 0
140 160 180 200 220 240 260 280 300 140 160 180 200 220 240 260 280 300
EP

Temperature /oC Temperature /oC


Fig. 7. Conversion of ethanol over studied catalysts.

9 It is generally accepted that both the microporous character and intrinsic acidity are the key factors
C

10 influencing the catalytic performance in the reactions involving small hydrocarbons. A quality of
AC

11 microporous character of zeolites ZSM-5 tested in MTH reaction [59] found the reflection in the
12 enhanced total light olefin selectivity. For ethanol dehydration it has been reported that the better
13 developed intrinsic acidity and micropore area are, providing confinement effect for adsorption of
14 small alkane or alcohol reactant molecules, the higher catalytic activity in this type of reaction [2].
15 Additionally, the development of mesopore surface area upon subsequent desilication of
16 dealuminated samples should find reflection in their stability and resistance against coke formation.
17 The lowest catalytic activity was recognized for zeolite n-L, i.e. zeolite characterized by the high
18 crystallinity and fully preserved microporous structure. From a catalytic point of view, the sites

21
ACCEPTED MANUSCRIPT
1 located in cancrinite cages are hardly accessible for reagent molecules and they do not deliver any
2 catalytic performance. However, the protonic sites located at the front of the boat shaped eight-
3 membered rings are available from 12-MR channels, thus they can offer a positive impact on catalytic
4 performance. In spite of the formation of the amorphous phase at the expense of the microporous
5 environment, the materials offering the most developed mesopore surface area were found to be
6 more effective catalysts. During desilication one-dimensional pore structure of zeolite L is believed to
7 be altered, especially in the enhancement of the population of Al-rich eight-membered rings

PT
8 responsible for the confinement effect. Subsequent dealumination process is thus more effective and
9 offers the significant changes in the acidity of the resulting materials. The acid sites previously

RI
10 engaged in hydrogen bonding become to be isolated species therefore easily reached by reagent
11 molecule. More developed mesopore surface area in the materials obtained via subsequent

SC
12 dealumination points to their predominance over the solely dealuminated ones. The most active
13 materials L_I-c and L-I-d showed the highest selectivity to ethylene reaching the level of 96-99 % of

U
14 conversion in the temperature range 260 – 300 oC (Fig. 7). The TOF values confirm that the highest
15 performance of active sites was observed for the alkaline/acid modified samples (Table 3). . Higher
AN
16 hydrocarbons were not produced in significant amounts (Fig. 8). High reaction temperature (280-300
o
17 C) favored intramolecular dehydration to ethylene, avoiding the formation of diethyl ether and
M

18 oligomers. It is generally accepted that ethanol dehydration involves the generation of the ethoxy
19 groups thus both Brønsted and Lewis acid sites may participate in the elimination mechanism. It can
D

20 be seen that during 35 h of the test at 280 oC for L_I-c sample an ethanol conversion was nearly
21 constant and reaching the level of 94 % (Fig. 9A). Thus, it could be concluded that the studied L_I-c
TE

22 catalyst presents relatively high stability under reaction conditions. When looking at selectivities (Fig.
23 9B) it is well seen that L_I-c material produced ethylene with selectivity above 95% for an overall
EP

24 time of the test. The zeolite HZSM-5 is widely considered as the catalyst in the dehydration of
25 ethanol to ethylene in the temperature range of 200 to 300 °C [2] reaching an ethanol conversion
C

26 level of 98% and 95% ethylene selectivity at 300 °C. However, high acidity ZSM-5 reduces its stability
27 due to the enhanced coke formation. Many approaches dedicated to an enhancement of coke
AC

28 stability has been developed. Zhang et al. [60] and Ramesh et al. [61] found that phosphorus
29 modified zeolite HZSM-5 allowed for maintaining high ethylene selectivity and increasing coking
30 resistance. Also, hierarchical HZSM-5 catalysts compared to conventional-sized HZSM-5 exhibited
31 high stability while reaching high ethylene selectivity at 100% ethanol conversion at 240-260 °C.
32 Herein we demonstrate that modification of zeolite L by alkaline/acid treatment can deliver ethylene
33 selectivity at the level of 98% at 100% ethanol conversion. The stability tests (Fig. 9) were also
34 performed on native zeolite n-L and zeolite HSM-5 of Si/Al=11.5 thus of the aluminium content as
35 much as possible similar to that in zeolites L studied.

22
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D

Fig. 8. The selectivities to (A) diethyl ether (DEE), (B) ethylene (C2=) and (C) higher hydrocarbons (C3+) over the most active
TE

catalysts obtained via I route.

1 Table 3. The turnover frequencies values (TOF ) for studied catalysts in different temperatures of the reaction.
EP

TOF /10-2 s-1


Sample
160°C 220°C 280°C
n-L 0.6 48.8 34.3
C

L_I-a 0.6 24.2 43.9


L_I-b 1.2 25.9 49.8
AC

L_I-c 3.4 75.0 114.1


L_I-d 3.5 79.1 115.2
L_II-a 0.4 18.9 62.2
L_II-c 0.4 14.1 46.2
L_II-d 0.4 21.4 81.3
L_III-b 1.0 23.7 42.5
L_III-d 0.4 14.6 29.0
L_IV-b 1.0 39.6 74.3
L_IV-c 0.6 42.7 81.3

23
ACCEPTED MANUSCRIPT

PT
RI
SC
o
Fig. 9. The conversion (A) and the selectivities (B, C and D) curves of stability tests performed at 280 C for n-L, L_I-c, and
HZSM-5 zeolites.

U
1 Indeed, for ZSM-5, in spite of the highest value of the conversion (99 %) among zeolites studied the
AN
2 selectivity for ethylene after the initial time of reaction was slightly above 70 %. A high strength of
3 protonic sites in ZSM-5 reduced its selectivity to the ethylene and enhanced the formation of C3+
4 moieties; the selectivity to C3+ products was maintained around 30 % through the whole stability test.
M

5 The native zeolite n-L was found to be the less active, however more selective to ethylene than ZSM-
6 5 especially in the first period of the test. For n-L zeolite as the selectivity to ethylene was decreased
D

7 the DEE was produced in higher amounts. Also, the higher hydrocarbons were not produced, only the
TE

8 negligible amount of acetaldehyde was formed. The modified zeolite L_I-c obtained by sequential
9 desilication and dealumination is promising catalyst for ethanol dehydration, since the selectivity to
10 ethylene was maintained on the level of 95% through the whole period of stability test.
EP

11 The spent catalyst L_I-c was investigated in the term of the coke precursors formation. The Raman
C

12 spectroscopy is a technique particularly dedicated to an examination of not only the extended


13 crystalline structures but also toward smaller entities with short-range order and molecular
AC

14 structures. Therefore this technique was used to investigate the phase composition of carbon
15 deposit in the most active catalysts after stability test reaction (Fig. 10). A broad but less intensive
16 band of the complex nature were found between 1000 and 1800 cm−1. After a numerical
17 deconvolution, the five individual Lorentzian components can be distinguished. Since the CH2 scissor
18 band at 1450 cm−1, characteristic of n-paraffins, is not observed in the spectrum, it is believed that
19 the detected hydrocarbons were alkyl polyaromatic species [62]. The vibrational features were
20 assigned to CC or ring stretches of polyaromatic hydrocarbons. Consequently, the bands at about
21 1310 cm−1 and 1375 cm−1 were classified as originating from the carbon sp2 in aromatic rings. The

24
ACCEPTED MANUSCRIPT
1 1485 cm−1 band is a common structural element in the five-membered-ring aromatic compound, thus
2 it was attributed to the in-phase C=C stretch of in an alkylated cyclopentadienyl species. The
3 1585 cm−1 band characterizes C=C stretches of conjugated olefins while the frequency of the
4 1650 cm−1 band points to the presence of isolated C=C double bonds (1640–1660 cm−1) associated
5 with disordered aromatic structures. For the polyaromatic compounds with a two-dimensional,
6 sheet-like topology the intensity in the 1600 to 1650 cm−1 region is significantly higher than that in
7 1300 to 1450 cm−1 region, thus in can be concluded that the polyaromatic hydrocarbons formed in

PT
8 spent L_I-c have a chain-like topology, rather than a two-dimensional, sheet-like topology. Schulz and
9 Wei [63] proposed a mechanism of coke formation wherein conjugated olefins undergo cyclization to

RI
10 form cyclopentadienyl species which subsequently form aromatic and, ultimately, polyaromatic
11 hydrocarbons. Spectroscopic detection, by UV Raman spectroscopy, of conjugated olefins,

SC
12 cyclopentadienyl species, and polyaromatic hydrocarbons were attained in zeolite H-MFI during the
13 MTH reaction [62]. The presence of the 1485 cm−1 band attributed to a cyclopentadienyl species in

U
14 spent catalyst L-I-c points to the role of cyclopentadienyl species in its coking during ethanol
15 dehydration. The slight decrease in the conversion to 96 % observed for L-I-c can be rationalized in
AN
16 the term of the formation of the coke precursors, still however in the amounts small enough to
17 preserve high activity and selectivity of the zeolitic catalyst.
M
D 85

75
14

13
85
TE 50
15

10
16

13
EP

1800 1600 1400 1200 1000


-1
Raman Shift (cm )
C

Fig. 10. Raman spectra of the spent catalyst L_I-c and the results of a fitting procedure.
AC

18 Conclusions

19 It has been demonstrated that for zeolite L chosen properly sequential dealumination/desilication
20 procedures offered the great mesoporosity development (up to 134 m2·g-1) with the preservation of
21 micropore volume and crystallinity. However, in the course of the dehydration of ethanol leading to
22 obtaining ethylene, the key parameter was the number and nature of the acidic Si(OH)Al. It has been
23 proven that the highest catalytic activity and selectivity to ethylene was obtained for the
24 alkaline/acid treated zeolites L accommodating solely the isolated Si(OH)Al groups, while the

25
ACCEPTED MANUSCRIPT
1 modification routes resulting in the increase in the number of the hydrogen bonded Si(OH)Al…O
2 species were associated with the reduction of catalytic activity. Moreover, the modified hierarchical
3 zeolites L turned out to be highly stable and resistant to the coke formation. After 35 hours of the
4 stability test the polyaromatic hydrocarbons, among them cyclopentadienyl species, were defined as
5 the precursor coke. Still, however, the amounts of the coke precursors were small enough to
6 preserve high activity and selectivity of the zeolitic catalyst.
7

PT
8
9 ACKNOWLEDGMENTS

RI
10 This work was financed by Grant No. 2014/13/D/ST5/02761 from the National Science Centre,
11 Poland. The 29Si MAS NMR studies were collected thanks to the H. Niewodniczański Institute Nuclear

SC
12 Physics PAS, in Cracow. K.G.-M. thanks for financial support from the National Science Centre,
13 Poland, Grant No. 2015/18/E/ST4/00191.

14 4. References
U
AN
15 [1] M. Maillard, S. Giorgio, M.-P. Pileni, Tuning the Size of Silver Nanodisks with Similar Aspect
16 Ratios:  Synthesis and Optical Properties, The Journal of Physical Chemistry B, 107 (2003) 2466-2470.
17 [2] K.A. Tarach, J. Tekla, W. Makowski, U. Filek, K. Mlekodaj, V. Girman, M. Choi, K. Gora-Marek,
18 Catalytic dehydration of ethanol over hierarchical ZSM-5 zeolites: studies of their acidity and porosity
M

19 properties, Catalysis Science & Technology, 6 (2016) 3568-3584.


20 [3] K. Tarach, K. Gora-Marek, J. Tekla, K. Brylewska, J. Datka, K. Mlekodaj, W. Makowski, M.C.
21 Igualada Lopez, J. Martinez Triguero, F. Rey, Catalytic cracking performance of alkaline-treated
D

22 zeolite Beta in the terms of acid sites properties and their accessibility, Journal of Catalysis, 312
23 (2014) 46-57.
TE

24 [4] D. Macina, Z. Piwowarska, K. Gora-Marek, K. Tarach, M. Rutkowska, V. Girman, A. Blachowski, L.


25 Chmielarz, SBA-15 loaded with iron by various methods as catalyst for DeNOx process, Materials
26 Research Bulletin, 78 (2016) 72-82.
EP

27 [5] H. Oikawa, Y. Shibata, K. Inazu, Y. Iwase, K. Murai, S. Hyodo, G. Kobayashi, T. Baba, Highly
28 selective conversion of ethene to propene over SAPO-34 as a solid acid catalyst, Applied Catalysis A:
29 General, 312 (2006) 181-185.
30 [6] W. Choopun, S. Jitkarnka, Catalytic activity and stability of HZSM-5 zeolite and hierarchical
C

31 uniform mesoporous MSU-SZSM-5 material during bio-ethanol dehydration, Journal of Cleaner


32 Production, 135 (2016) 368-378.
AC

33 [7] I.S. Yakovleva, S.P. Banzaraktsaeva, E.V. Ovchinnikova, V.A. Chumachenko, L.A. Isupova, Catalytic
34 dehydration of bioethanol to ethylene, Catalysis in Industry, 8 (2016) 152-167.
35 [8] K. Gora-Marek, IR studies of the transformation of formaldehyde and methanol on Co-ferrierites,
36 Microporous and Mesoporous Materials, 145 (2011) 93-97.
37 [9] A. Stepniewski, M. Radon, K. Gora-Marek, E. Broclawik, Ammonia-modified Co(ii) sites in zeolites:
38 spin and electron density redistribution through the Co(II)-NO bond, Physical chemistry chemical
39 physics : PCCP, 18 (2016) 3716-3729.
40 [10] P. Pietrzyk, C. Dujardin, K. Gora-Marek, P. Granger, Z. Sojka, Spectroscopic IR, EPR, and operando
41 DRIFT insights into surface reaction pathways of selective reduction of NO by propene over the Co-
42 BEA zeolite, Physical Chemistry Chemical Physics, 14 (2012) 2203-2215.
43 [11] R.C. Hansford, J.W. Ward, The nature of active sites on zeolites. VII. Relative activities of
44 crystalline and amorphous alumino-silicates, Journal of Catalysis, 13 (1969) 316-320.

26
ACCEPTED MANUSCRIPT
1 [12] Z.T. Lalowicz, G. Stoch, A. Birczynski, M. Punkkinen, M. Krzystyniak, K. Gora-Marek, J. Datka,
2 Dynamics of hydroxyl deuterons and bonded water molecules in NaDY(0.8) zeolite as studied by
3 means of deuteron NMR spectroscopy and relaxation, Solid State Nuclear Magnetic Resonance, 37
4 (2010) 91-100.
5 [13] T.F. Degnan Jr, The implications of the fundamentals of shape selectivity for the development of
6 catalysts for the petroleum and petrochemical industries, Journal of Catalysis, 216 (2003) 32-46.
7 [14] A.M. W, K. Jacek, Solid-state NMR studies of the shape-selective catalytic conversion of
8 methanol into gasoline on zeolite ZSM-5.
9 [15] C. de las Pozas, R. Lopez-Cordero, J.A. Gonzalez-Morales, N. Travieso, R. Roque-Malherbe, Effect
10 of pore diameter and acid strength in ethanol dehydration on molecular sieves, Journal of Molecular

PT
11 Catalysis, 83 (1993) 145-156.
12 [16] D. Macina, Z. Piwowarska, K. Tarach, K. Gora-Marek, J. Ryczkowski, L. Chmielarz, Mesoporous
13 silica materials modified with alumina polycations as catalysts for the synthesis of dimethyl ether

RI
14 from methanol, Materials Research Bulletin, 74 (2016) 425-435.
15 [17] A.M. Szymocha, A. Birczynski, Z.T. Lalowicz, G. Stoch, M. Krzystyniak, K. Gora-Marek, Water
16 confinement in faujasite cages: a deuteron NMR investigation in a wide temperature range. 1. Low

SC
17 temperature spectra, The journal of physical chemistry. A, 118 (2014) 5359-5370.
18 [18] J. Datka, B. Gil, T. Domagala, K. Gora-Marek, Homogeneous OH groups in dealuminated HY
19 zeolite studied by IR spectroscopy, Microporous and Mesoporous Materials, 47 (2001) 61-66.
20 [19] C.J. Van Oers, K. Gora-Marek, B. Prelot, J. Datka, V. Meynen, P. Cool, Demonstrating the Benefits

U
21 and Pitfalls of Various Acidity Characterization Techniques by a Case Study on Bimodal
22 Aluminosilicates, Langmuir, 30 (2014) 1880-1887.
AN
23 [20] M. Moreno-Gonzalez, T. Blasco, K. Gora-Marek, A.E. Palomares, A. Corma, Study of propane
24 oxidation on Cu-zeolite catalysts by in-situ EPR and IR spectroscopies, Catalysis Today, 227 (2014)
25 123-129.
M

26 [21] A.M. Szymocha, Z.T. Lalowicz, A. Birczynski, M. Krzystyniak, G. Stoch, K. Gora-Marek, Water
27 confinement in faujasite cages: a deuteron NMR investigation in a wide temperature range. 2.
28 Spectra and relaxation at high temperature, The journal of physical chemistry. A, 118 (2014) 5371-
D

29 5380.
30 [22] C.J. Van Oers, K. Gora-Marek, K. Sadowska, M. Mertens, V. Meynen, J. Datka, P. Cool, In situ IR
31 spectroscopic study to reveal the impact of the synthesis conditions of zeolite beta nanoparticles on
TE

32 the acidic properties of the resulting zeolite, Chemical Engineering Journal, 237 (2014) 372-379.
33 [23] K. Sadowska, K. Gora-Marek, M. Drozdek, P. Kustrowski, J. Datka, J. Martinez Triguero, F. Rey,
34 Desilication of highly siliceous zeolite ZSM-5 with NaOH and NaOH/tetrabutylamine hydroxide,
EP

35 Microporous and Mesoporous Materials, 168 (2013) 195-205.


36 [24] K. Gora-Marek, K.A. Tarach, Z. Piwowarska, M. Laniecki, L. Chmielarz, Ag-loaded zeolites Y and
37 USY as catalysts for selective ammonia oxidation, Catalysis Science &amp; Technology, 6 (2016)
38 1651-1660.
C

39 [25] P. Pietrzyk, K. Gora-Marek, Paramagnetic dioxovanadium(iv) molecules inside the channels of


40 zeolite BEA - EPR screening of VO2 reactivity toward small gas-phase molecules, Physical chemistry
AC

41 chemical physics : PCCP, 18 (2016) 9490-9496.


42 [26] D. Majda, K. Tarach, K. Gora-Marek, A. Michalik-Zym, S.D. Napruszewska, M. Zimowska, E.M.
43 Serwicka, Thermoporosimetry of n-alkanes for characterization of mesoporous SBA-15 silicas -
44 Towards deeper understanding the effect of the probe liquid nature, Microporous and Mesoporous
45 Materials, 226 (2016) 25-33.
46 [27] P. Pietrzyk, K. Gora-Marek, Paramagnetic dioxovanadium(IV) molecules inside the channels of
47 zeolite BEA - EPR screening of VO2 reactivity toward small gas-phase molecules, Physical Chemistry
48 Chemical Physics, 18 (2016) 9490-9496.
49 [28] K. Sadowska, K. Gora-Marek, J. Datka, Hierarchic zeolites studied by IR spectroscopy: Acid
50 properties of zeolite ZSM-5 desilicated with NaOH and NaOH/tetrabutylamine hydroxide, Vibrational
51 Spectroscopy, 63 (2012) 418-425.

27
ACCEPTED MANUSCRIPT
1 [29] F.F. Madeira, N.S. Gnep, P. Magnoux, S. Maury, N. Cadran, Ethanol transformation over HFAU,
2 HBEA and HMFI zeolites presenting similar Brønsted acidity, Applied Catalysis A: General, 367 (2009)
3 39-46.
4 [30] D.W. Breck, A. N.A., US Patent 2,711,565, 1958.
5 [31] P. Pichat, C. Franco-Parra, D. Barthomeuf, Infra-red structural study of various type L zeolites,
6 Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases, 71
7 (1975) 991-996.
8 [32] R.M. Barrer, H. Vllliger, The crystal structure of the synthetic zeolite l, Zeitschrift fur
9 Kristallographie - New Crystal Structures, 1969, pp. 352-370.
10 [33] P. Bartl, W.F. Hölderich, A study of the dealumination methods for zeolite L, Microporous and

PT
11 Mesoporous Materials, 38 (2000) 279-286.
12 [34] S.S. Hassani, F. Salehirad, H.R. Aghabozorg, Z. Sobat, Synthesis and morphology of nanosized
13 zeolite L, Crystal Research and Technology, 45 (2010) 183-187.

RI
14 [35] J.M. Newsam, M.T. Melchior, H. Malone, Full profile analysis of the 29Si NMR spectra of LTL-
15 framework zeolites, Solid State Ionics, 26 (1988) 125-131.
16 [36] W.F. Hölderich, H.v. Bekkum, E.M.F.a.J.C.J. H. van Bekkum, Chapter 16 “Zeolites in Organic

SC
17 Syntheses”, Studies in Surface Science and Catalysis, Elsevier1991, pp. 631-726.
18 [37] R.E. Jentoft, M. Tsapatsis, M.E. Davis, B.C. Gates, Platinum Clusters Supported in Zeolite LTL:
19 Influence of Catalyst Morphology on Performance inn-Hexane Reforming, Journal of Catalysis, 179
20 (1998) 565-580.

U
21 [38] H. Chen, J. Wydra, X. Zhang, P.-S. Lee, Z. Wang, W. Fan, M. Tsapatsis, Hydrothermal Synthesis of
22 Zeolites with Three-Dimensionally Ordered Mesoporous-Imprinted Structure, Journal of the
AN
23 American Chemical Society, 133 (2011) 12390-12393.
24 [39] Q. Huo, T. Dou, Z. Zhao, H. Pan, Synthesis and application of a novel mesoporous zeolite L in the
25 catalyst for the HDS of FCC gasoline, Applied Catalysis A: General, 381 (2010) 101-108.
M

26 [40] D. Verboekend, T.C. Keller, M. Milina, R. Hauert, J. Pérez-Ramírez, Hierarchy Brings Function:
27 Mesoporous Clinoptilolite and L Zeolite Catalysts Synthesized by Tandem Acid–Base Treatments,
28 Chemistry of Materials, 25 (2013) 1947-1959.
D

29 [41] J.C. Groen, L.A.A. Peffer, J.A. Moulijn, J. Pérez-Ramírez, Mechanism of hierarchical porosity
30 development in MFI zeolites by desilication: The role of aluminium as a pore-directing agent,
31 Chemistry - A European Journal, 11 (2005) 4983-4994.
TE

32 [42] K. Góra-Marek, M. Derewiński, P. Sarv, J. Datka, IR and NMR studies of mesoporous alumina and
33 related aluminosilicates, Catalysis Today, 101 (2005) 131-138.
34 [43] K. Sadowska, K. Góra-Marek, J. Datka, Hierarchic zeolites studied by IR spectroscopy: Acid
EP

35 properties of zeolite ZSM-5 desilicated with NaOH and NaOH/tetrabutylamine hydroxide, Vibrational
36 Spectroscopy, 63 (2012) 418-425.
37 [44] J. Datka, K. Gora-Marek, IR studies of the formation of ammonia dimers in zeolites TON, Catalysis
38 Today, 114 (2006) 205-210.
C

39 [45] S.M. Csicsery, D.A. Hickson, Selectivities of acid-catalyzed reactions over silica-alumina and
40 molecular sieve catalysts, Journal of Catalysis, 19 (1970) 386-393.
AC

41 [46] D. Verboekend, J. Pérez-Ramírez, Desilication Mechanism Revisited: Highly Mesoporous All-Silica


42 Zeolites Enabled Through Pore-Directing Agents, Chemistry – A European Journal, 17 (2011) 1137-
43 1147.
44 [47] M.M.J. Treacy, J.B. Higgins, R.v. Ballmoos, Preface, Zeolites, 16 (1996) 327-328.
45 [48] P. Huifang, L. Xingyu, W. Guoyou, Y. Hailiang, P. Xinghong, H. Quan, Y. Pingxiang, Synthesis of La-
46 modified ultra stable zeolite L and its application to catalytic cracking catalyst, Petroleum Science, 4
47 (2007) 75-81.
48 [49] M. Hunger, J. Karger, H. Pfeifer, J. Caro, B. Zibrowius, M. Bulow, R. Mostowicz, Investigation of
49 internal silanol groups as structural defects in ZSM-5-type zeolites, Journal of the Chemical Society,
50 Faraday Transactions 1: Physical Chemistry in Condensed Phases, 83 (1987) 3459-3468.

28
ACCEPTED MANUSCRIPT
1 [50] J. Kornatowski, W.H. Baur, G. Pieper, M. Rozwadowski, W. Schmitz, A. Cichowlas, Dealumination
2 of large crystals of zeolite ZMS-5 by various methods, Journal of the Chemical Society, Faraday
3 Transactions, 88 (1992) 1339-1343.
4 [51] D. Zhai, L. Zhao, Y. Liu, J. Xu, B. Shen, J. Gao, Dissolution and Absorption: A Molecular Mechanism
5 of Mesopore Formation in Alkaline Treatment of Zeolite, Chemistry of Materials, 27 (2015) 67-74.
6 [52] K.A. Tarach, J. Tekla, W. Makowski, U. Filek, K. Mlekodaj, V. Girman, M. Choi, K. Gora-Marek,
7 Catalytic dehydration of ethanol over hierarchical ZSM-5 zeolites: studies of their acidity and porosity
8 properties, Catalysis Science &amp; Technology, 6 (2016) 3568-3584.
9 [53] V.L. Zholobenko, L.M. Kustov, V.Y. Borovkov, V.B. Kazansky, A new type of acidic hydroxyl groups
10 in ZSM-5 zeolite and in mordenite according to diffuse reflectance i.r. spectroscopy, Zeolites, 8 (1988)

PT
11 175-178.
12 [54] E. Brunner, K. Beck, M. Koch, L. Heeribout, H.G. Karge, Verification and quantitative
13 determination of a new type of Brønsted acid sites in H-ZSM-5 by 1H magic-angle spinning nuclear

RI
14 magnetic resonance spectroscopy, Microporous Materials, 3 (1995) 395-399.
15 [55] T. Takaishi, Ordered distribution of aluminium or gallium atoms in zeolite L, Journal of the
16 Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases, 84 (1988) 2967-

SC
17 2977.
18 [56] G. Turnes Palomino, C. Otero Arean, F. Geobaldo, G. Ricchiardi, S. Bordiga, A. Zecchina, FTIR
19 study of CO adsorbed at low temperature on zeolite L Evidence for an ordered distribution of
20 aluminium atoms, Journal of the Chemical Society, Faraday Transactions, 93 (1997) 189-191.

U
21 [57] M. John, K. Alexopoulos, M.-F. Reyniers, G.B. Marin, First-Principles Kinetic Study on the Effect of
22 the Zeolite Framework on 1-Butanol Dehydration, ACS Catalysis, 6 (2016) 4081-4094.
AN
23 [58] H. Chiang, A. Bhan, Catalytic consequences of hydroxyl group location on the rate and
24 mechanism of parallel dehydration reactions of ethanol over acidic zeolites, Journal of Catalysis, 271
25 (2010) 251-261.
M

26 [59] J. Kim, M. Choi, R. Ryoo, Effect of mesoporosity against the deactivation of MFI zeolite catalyst
27 during the methanol-to-hydrocarbon conversion process, Journal of Catalysis, 269 (2010) 219-228.
28 [60] J.W. Ward, The nature of active sites on zeolites. VIII. Rare earth Y zeolite, Journal of Catalysis,
D

29 13 (1969) 321-327.
30 [61] K. Ramesh, L.M. Hui, Y.-F. Han, A. Borgna, Structure and reactivity of phosphorous modified H-
31 ZSM-5 catalysts for ethanol dehydration, Catalysis Communications, 10 (2009) 567-571.
TE

32 [62] Y.T. Chua, P.C. Stair, An ultraviolet Raman spectroscopic study of coke formation in methanol to
33 hydrocarbons conversion over zeolite H-MFI, Journal of Catalysis, 213 (2003) 39-46.
34 [63] H. Schulz, M. Wei, Deactivation and thermal regeneration of zeolite HZSM-5 for methanol
EP

35 conversion at low temperature (260–290°C), Microporous and Mesoporous Materials, 29 (1999) 205-
36 218.

37
C
AC

29
ACCEPTED MANUSCRIPT
Hierarchical L zeolites as highly stable catalysts in ethanol dehydration.

The type of acidic Si(OH)Al groups determine the catalytic activity.

For L zeolite sequential demetallation offer the mesoporosity up to 134 m2·g-1.

PT
RI
U SC
AN
M
D
TE
C EP
AC

Vous aimerez peut-être aussi