Preparation of Mesoporous Mordenite For The Hydroisomerization of N-Hexane

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Catalysis Communications 125 (2019) 21–25

Contents lists available at ScienceDirect

Catalysis Communications
journal homepage: www.elsevier.com/locate/catcom

Short communication

Preparation of mesoporous mordenite for the hydroisomerization of n- T


hexane

Xinqing Lu, Yuping Guo, Chunhui Xu, Rui Ma , Xue Wang, Ningwei Wang, Yanghe Fu,

Weidong Zhu
Key Laboratory of the Ministry of Education for Advanced Catalysis Materials, Institute of Physical Chemistry, Zhejiang Normal University, 321004 Jinhua, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: A mesoporous mordenite (MOR) zeolite was prepared from a commercially available MOR sample by a se-
Mesoporous zeolite quential acid-alkaline-acid post-treatment. The latter was used as support of Pt to prepare a bifunctional Pt/H-
Bifunctional catalyst MOR catalyst for the hydroisomerization of n-hexane. The results showed that the available mesopores in the
Mordenite catalyst shorten the diffusion length of reactants and of the produced branched hexanes inside the micropores. In
Desilication
addition, the dispersion of supported Pt in the zeolite was increased and an enhancement in the conversion of n-
Hydroisomerization
hexane and at the same time the reduction of the non-selective cracking of hexane isomers and of the formation
of coke (catalyst deactivation) was observed.

1. Introduction sites in zeolites could be enhanced by introducing mesopores in the


zeolite pore structure by either direct syntheses with soft and hard
The hydroisomerization of n-paraffins into isoparaffins plays an templates [9,10] or post-treatments with acid [11], alkaline [12] and
important role in improving the quality of gasoline by increasing the NH4F [13]. Alkaline-derived desilication has been used as an efficient
octane number. The commercial process consists of (de)hydrogenation strategy for the generation of mesopores in zeolite secondary crystals
on metallic Pt and skeletal isomerization on acid sites typically cata- [14–21]. However, the Si/Al molar ratios in zeolites should be in an
lyzed by bifunctional Pt/H-MOR catalysts, which is generally accom- optimal range with the purpose to preserve the zeolite's crystallinity
panied with cracking reactions [1]. The balance between (de)hydro- and acidity [22–24]. If these ratios become too low, no mesoporosity is
genation and skeletal isomerization plays a determinant role on the formed, while if these ratios become too high, then excessive silicon
activity and selectivity for branched products in this isomerization dissolution results in the collapse of the zeolite structure. Acid treat-
process [2,3]. Additionally, the product selectivity is significantly af- ment was used to increase the Si/Al ratio of MOR into its optimal range
fected by the proximity between metallic Pt and acid sites [4,5]. MOR via dealumination, making mesopores successfully formed by an alka-
has two types of channels, 12-membered-ring channels with an opening line treatment [23]. van Laak et al. [24] found an improved reactivity
size of 0.65 nm × 0.70 nm, and 8-membered-ring channels with an in the liquid-phase alkylation of benzene with propylene to cumene
opening size of 0.26 nm × 0.57 nm. Therefore, in general, MOR is re- over the mesoporous MOR catalyst prepared by a sequential acid-al-
garded as one-dimensional pore system, because the 8-membered-ring kaline-acid post-treatment. A similar strategy was also applied for
channels are too small for most molecules to enter, inducing one-di- preparing a mesoporous MOR-type titanosilicate that showed re-
mensional diffusion, i.e., molecules cannot overtake each other in the markably improved catalytic properties in the hydroxylation of toluene
channels so a molecule located at the pore mouth blocks the entrance or and the ammoximation of cyclohexanone [25]. Most recently, Pastvova
escape of other molecules [6,7]. Hence, the molecular transport in the et al. [26] applied a sequential acid-alkaline-acid post-treatment to
micropores is strongly hindered by this one-dimensional diffusion in generate the mesoporosity in MOR and then used this mesoporous
comparison with the 2-D or 3-D pore systems [6–8]. Additionally, the zeolite as a support to prepare a Pt/H-MOR catalyst for the hydro-
lodgment of extra-framework species in the MOR channels can sig- isomerization of n-hexane, showing a significantly improved catalytic
nificantly restrict the accessibility to adsorbate and/or reactant mole- activity due to the simultaneous enhancement of the accessibility of the
cules [7]. acid sites and of molecular transport. However, less work dealt with the
It has been repeatedly demonstrated that the accessibility of the acid effects of the introduced mesopores on the dispersion of Pt in the zeolite


Corresponding authors.
E-mail addresses: mr@zjnu.cn (R. Ma), weidongzhu@zjnu.cn (W. Zhu).

https://doi.org/10.1016/j.catcom.2019.03.017
Received 16 January 2019; Received in revised form 18 March 2019; Accepted 19 March 2019
Available online 20 March 2019
1566-7367/ © 2019 Elsevier B.V. All rights reserved.
X. Lu, et al. Catalysis Communications 125 (2019) 21–25

and its derived catalytic selectivity and stability. 2.2. Catalyst characterization
In the present work, a rich mesoporosity was introduced into MOR
secondary crystals via a sequential acid-alkaline-acid post-treatment. A The powder X-ray diffraction (XRD) patterns of the solid catalysts
comparative study was conducted on the catalytic properties of Pt/H- prepared were recorded by a Philips PW3040/60 powder dif-
MOR catalysts prepared using commercially available, acid post- fractometer using Ni-filtered Cu Kα radiation (λ = 0.1541 nm) in the
treated, and sequential acid-alkaline-acid post-treated MOR samples as range of 5–50° with a scanning rate of 2° min−1. The zeolite crystal
supports for the hydroisomerization of n-hexane in order to investigate morphology was examined by a Hitachi S-4800 scanning electron mi-
the effects of the introduced mesopores in MOR secondary crystals on croscope (SEM). Transmission electron microscopy (TEM) for Pt par-
the derived catalytic activity, selectivity and stability. ticle size and mesoporosity characterization was carried out on a JEOL
JEM-2100 microscope after the sample was deposited onto a holey
carbon foil supported on a copper grid. The textural properties of the
2. Experimental samples were determined by N2 adsorption-desorption at 77 K using a
Micromeritics ASAP 2020 instrument after the samples were degassed
2.1. Catalyst preparation in vacuum at 573 K for 6 h. The specific surface area of the investigated
samples was calculated using the Brunauer-Emmett-Teller (BET)
2.1.1. Acid and alkaline post-treatments method in the relative pressure range of p/p0 = 0.05–0.35. The mi-
The MOR zeolite with a Si/Al molar ratio of 10 in an ammonium cropore volume and the external specific surface area were determined
form was purchased from Zeolyst (CBV 21A), denoted as NH4-MOR-p. by the t-plot method, and the total pore volume was determined by
The mesoporous MOR was obtained by a sequential acid-alkaline-acid converting the amount adsorbed at a relative pressure of 0.995 to the
treatment, similar to the procedure described elsewhere [23,24]. The volume of liquid nitrogen. The mesopore volume was computed by
as-received Zeolyst MOR was calcined at 773 K for 3 h with a heating subtracting the volume of micropores from the total pore volume.
rate of 1 K min−1 to get a proton-form MOR, denoted as H-MOR-p. The bulk Si/Al ratio and the amount of Pt loading in the catalyst
Afterwards, H-MOR-p was treated with 3 M HNO3 aqueous solution were determined by inductively coupled plasma-atomic emission
with a solid-to-liquid weight ratio of 1:10 (w/w) at 373 K for 1 h under spectrometry (ICP-AES) on a Thermo IRIS Intrepid II XSP after the
stirring. The acid-treated product was filtered, washed with hot-deio- samples were dissolved in liquated NaOH. The 27Al MAS NMR spectra
nized water until the pH of the filtrate became ~ 7. Then, the obtained were collected on a Varian VNMRS-400 MB NMR spectrometer. The
filter cake was dried at 383 K overnight and calcined at 773 K for 3 h spectra were collected at a frequency of 104.18 MHz, a spinning rate of
with a heating rate of 1 K min−1 to get a partially dealuminated MOR 9.0 kHz and a recycling delay of 4 s. KAl(SO4)2·12H2O was used as re-
(denoted as H-MOR-p-a). The alkaline treatment was carried out over ference for chemical shift.
H-MOR-p-a in 0.2 M NaOH aqueous solution with a solid-to-liquid The amount of the acid sites was determined by temperature-pro-
weight ratio of 1:30 (w/w) under stirring at 343 K for 5 min. The solid- grammed desorption of NH3 (NH3-TPD) on a Micromeritics AutoChem
liquid mixture was subsequently centrifuged and decanted of the liquid. II 2920 instrument equipped with a thermal conductivity detector
The alkaline-treated product was washed with hot deionized water until (TCD). Typically, 80 mg of the sample was pretreated in flowing helium
the pH of the filtrate became ~ 7, and then dried at 383 K overnight and (25 mL min−1) at 823 K for 1 h. The adsorption of NH3 was carried out
calcined at 773 K for 3 h with a heating rate of 1 K min−1. The calcined at 393 K for 15 min, and then weakly chemisorbed NH3 was purged by
sample was ion-exchanged with 1 M NH4NO3 aqueous solution with a flowing helium (25 mL min−1) at the same temperature for 30 min. To
solid-liquid weight ratio of 1:12 (w/w) at 353 K for 12 h. To ensure the ensure the complete removal of weakly chemisorbed NH3, the purging
complete exchange of Na+, this procedure was repeated three times, process was repeated three times. The NH3-TPD profile was then re-
and the obtained solid sample was dried at 383 K overnight and cal- corded from 393 to 973 K with a heating rate of 15 K min−1 using He as
cined at 773 K for 3 h with a heating rate of 1 K min−1 to get a proton- carrier gas. The signal of desorbed NH3 was measured by TCD detector.
form MOR (denoted as H-MOR-p-a-a). Finally, H-MOR-p-a-a was wa-
shed with 0.1 M HNO3 aqueous solution with a solid-liquid weight ratio 2.3. Hydroisomerization of n-hexane
of 1:20 (w/w) at 323 K for 15 min under stirring, and the obtained solid
sample was then dried at 383 K overnight and calcined at 773 K for 3 h The hydroisomerization of n-hexane was carried out in a fixed-bed
with a heating rate of 1 K min−1 to get an acid-alkaline-acid treated reactor (6 mm inner diameter) at 1 atm pressure. 0.2 g of Pt/H-MOR
MOR (denoted as H-MOR-p-a-a-a). The flow diagram for the separate catalyst was loaded into the reactor and in-situ activated in 80 vol% H2
synthesis steps, including sample designations for the as-received Zeo- /20 vol% N2 gas mixture at 673 K for 2 h (total flow rate of
lyst MOR to the sequential acid-alkaline-acid post-treated MOR, is 10 mL min−1). After the reactor temperature was reduced to the reac-
presented in Scheme 1. tion temperature of 513 K, n-hexane was introduced into the reactor by
a high-pressure constant flow pump. The weight hourly space velocity
(WHSV) of n-hexane and the H2/n-hexane molar ratio for all the ex-
2.1.2. Pt/H-MOR preparation periments conducted were maintained at 2 h−1 and 6, respectively. The
Pt/H-MOR catalysts were prepared by the impregnation method as reaction products were analyzed by a gas chromatograph (Agilent
follows: a proton-form MOR sample (1 g) was impregnated with a Pt 6820) equipped with a 50 m OV-101 capillary column and a flame io-
(NH3)4Cl2 aqueous solution (0.5 mg g−1) of 2.5 mL initially to become a nization detector (FID).
slurry, followed by dropwise addition of the remaining Pt(NH3)4Cl2
aqueous solution of 7.5 mL into the slurry under vigorous stirring at 3. Results and discussion
308 K for 24 h. Then the final slurry was filtered, washed with hot-
deionized water until the complete removal of chloride ions and dried 3.1. Catalyst characterization
at 383 K overnight. Subsequently, the dried sample (1 g) was loaded in
a tubular reactor (5 cm inner diameter) and heated in flowing oxygen at As shown in Fig. S1 (ESI), all the investigated catalysts showed the
the flow rate of 500 mL min−1 at 673 K for 2 h with a heating rate of powder XRD characteristic patterns of MOR, indicating that the crystal
0.5 K min−1. The resultant catalysts with nominal loading of 0.5 wt% Pt structures are well preserved after the different post-treatments of the
and the various zeolite treated samples (see Section 2.1.1) are named as zeolite applied. Both Pt/H-MOR-p and Pt/H-MOR-p-a show a type-I N2
Pt/H-MOR-p, Pt/H-MOR-p-a, and Pt/H-MOR-p-a-a-a. adsorption-desorption isotherm (Fig. S2A), which is a characteristic of
microporosity. The acid-derived dealumination makes the micropore

22
X. Lu, et al. Catalysis Communications 125 (2019) 21–25

Scheme 1. Flow diagram of the different post-treatment steps, starting from the commercial zeolite.

Table 1
Textural properties and chemical composition of the catalysts investigated.
Catalysts SBET (m2 g−1) Vmicroa (cm3 g−1) Sexta (m2 g−1) Vtotal (p/p0 = 0.995) (cm3 g−1) Vmesob (cm3 g−1) Si/Alc Pt loadedc (wt%)

Pt/H-MOR-p 419 0.17 39 0.25 0.08 10 0.45


Pt/H-MOR-p-a 452 0.18 61 0.27 0.09 28 0.43
Pt/H-MOR-p-a-a-a 492 0.17 126 0.46 0.29 15 0.46

a
Determined from the measured N2 adsorption-desorption isotherms at 77 K using the t-plot method.
b
Vmeso = Vtotal–Vmicro.
c
Determined by the ICP-AES analysis.

volume and the external specific surface area sharply from


0.08 cm3 g−1 and 39 m2 g−1 for Pt/H-MOR-p to 0.29 cm3 g−1 and 126
m2 g−1 for Pt/H-MOR-p-a-a-a, respectively, while the micropore vo-
lume almost remains unchanged. Hence, these results confirm the va-
lidity of the sequential acid-alkaline-acid post-treatment as an effective
method to create mesoporosity in MOR crystals, while the micro-
porosity is preserved. Furthermore, the pore size distribution profiles
determined by the Barrett-Joyner-Halenda (BJH) method from the ad-
sorption branch, shown in Fig. S2B, indicate that Pt/H-MOR-p-a-a-a
possesses mesoporosity centered at ca. 10 nm, while Pt/H-MOR-p and
Pt/H-MOR-p-a are purely microporous. Additionally, the Si/Al molar
ratio is decreased from 28 for Pt/H-MOR-p-a to 15 for Pt/H-MOR-p-a-a-
a, respectively, due to the desilication (Table 1).
To elucidate the effects of dealumination and desilication of zeolites
on their acidity, 27Al MAS NMR spectra and NH3-TPD profiles were
measured. In the 27Al MAS NMR spectra (Fig. S3), Pt/H-MOR-p shows
only a resonance at 58 ppm ascribed to tetrahedral Al, the so-called
Fig. 1. NH3-TPD profiles of Pt/H-MOR-p (a), Pt/H-MOR-p-a (b), and Pt/H- framework Al species. The dealumination can develop extra-framework
MOR-p-a-a-a (c). Al species showing a new resonance at 0 ppm in the spectrum of Pt/H-
MOR-p-a, while these extra-framework Al species can be effectively
volume and the specific surface area slightly increased from removed by the alkaline-acid post-treatment, as the resonance at 0 ppm
0.17 cm3 g−1 and 419 m2 g−1 for Pt/H-MOR-p to 0.18 cm3 g−1 and 452 in the spectrum of Pt/H-MOR-p-a-a-a is almost absent.
m2 g−1 for Pt/H-MOR-p-a, respectively (Table 1) due to the enlarge- Fig. 1 shows the NH3-TPD profiles recorded after removing the
ment in the micropore size after the removal of Al in the framework weakly chemisorbed NH3 species on the investigated catalysts. The low-
[24,25]. Moreover, the Si/Al molar ratio is increased from 10 for Pt/H- temperature (400–525 K) and high-temperature (525–973 K) deso-
MOR-p to 28 for Pt/H-MOR-p-a, where the latter is in the optimum rption peaks are ascribed to the desorption of NH3 from the weak and
range of Si/Al molar ratio in the zeolite for the subsequent alkaline strong acid sites of MOR, respectively [26,27]. As shown in Fig. 1, the
post-treatment [24]. A further alkaline-acid post-treatment changes the strong acid sites, which are mainly considered as the catalytic sites in
isotherm to the combined one of type-I and type-IV with a well pro- the hydroisomerization, have been well preserved during the deal-
nounced hysteresis loop. The latter indicates the generation of meso- umination and desilication post-treatments. Compared to the high-
pores in the zeolite crystals (Fig. S2A). As shown in Table 1, the se- temperature desorption peak, the low-temperature peak was more ob-
quential acid-alkaline-acid post-treatment increases the mesopore viously affected by the variation of the Si/Al ratio (Table 1 and Fig. 1).

23
X. Lu, et al. Catalysis Communications 125 (2019) 21–25

The concentration of weak acid sites is in the order: Pt/H-MOR-p > Pt/
H-MOR-p-a-a-a > Pt/H-MOR-p-a, which is in accordance with the Al
amount in the zeolite. Additionally, the presence of a shoulder peak at
higher temperatures in the Pt/H-MOR-p-a indicates that the acid
strength is increased after dealumination due to the formation of extra-
framework Al species acting as Lewis acid sites [28].
As shown in Table 1, the Pt amounts in all the catalysts, determined
by the ICP-AES analysis, are within their nominal values of 0.5 wt%,
indicating the efficiency of the impregnation method applied for pre-
paring the catalysts. The SEM image of H-MOR-p shows that MOR
particles are aggregates of nanocrystals (Fig. S4a), while TEM images
reveal the mesoporosity in Pt/H-MOR-p-a-a-a and the distribution of
platinum oxide particle size in all catalysts (Figs. S4b-d). Pt/H-MOR-p
and Pt/H-MOR-p-a both show opaque crystals with some dark spots,
corresponding to the presence of platinum oxide particles located on
the surface of the MOR crystals (Figs. S3b and c). In contrast, the TEM
image of Pt/H-MOR-p-a-a-a shows many white spots with a diameter of Fig. 3. Distribution of products and the yield of total branched hexanes in the
ca. 10 nm (Fig. S4d), indicating the presence of mesopores in the MOR hydroisomerization of n-hexane over Pt/H-MOR-p (a), Pt/H-MOR-p-a (b) and
crystals, which is in accordance with the N2 adsorption-desorption Pt/H-MOR-p-a-a-a (c) in the first 60 min of reaction. The reaction conditions are
characterization results (Fig. S2). Moreover, the created mesopores are the same as those presented in Fig. 2. Others stands for C1-C5 products from the
indeed embedded in the crystals as shown in the TEM images of Fig. S5. cracking reaction.
The mean diameter of platinum oxide particles decreases from ca.
3.2 nm in Pt/H-MOR-p and ca. 3.0 nm in Pt/H-MOR-p-a to ca. 1.8 nm in Lewis acid sites (Fig. 2), and the cracking products are pronounced as
Pt/H-MOR-p-a-a-a, implying that the presence of mesopores in the shown in Fig. 3. It is thus illustrated that Pt/H-MOR-p-a-a-a shows the
zeolite enhanced the dispersion of Pt. highest conversion of n-hexane and yield of hexane isomers, with a
stable activity and at the same time negligible cracking products
3.2. Hydroisomerization of n-hexane (Figs. 2 and 3). The enhanced catalytic activity must be related to the
introduction of mesoporosity with the preservation of the zeolite fra-
The catalytic hydroisomerization of n-hexane consists of the dehy- mework crystallinity, microporosity and acidity. The introduced me-
drogenation of n-hexane to n-hexene on metallic Pt, the skeletal iso- sopores in the MOR by the sequential acid-alkaline-acid post-treatment
merization of n-hexene to monobranched methylpentenes and di- can shorten the diffusion length of the reactants and of the produced
branched dimethylbutenes on acid sites, and the hydrogenation of the branched hexanes inside the micropores [26], affect the proximity be-
branched alkene intermediates to the corresponding branched alkanes tween metallic Pt and acid sites [4,5], and enhance the dispersion of Pt
on metallic Pt [26]. In addition, the hydroisomerization of n-hexane to in the zeolite. All these factors eventually lead to the increase of con-
isohexane is usually accompanied with the cracking reaction and coke version of n-hexane and reduction of the non-selective cracking of
formation [1]. As shown in Figs. 2 and 3, Pt/H-MOR-p-a possesses a hexane isomers and the formation of coke. For all the investigated
higher conversion of n-hexane and a higher yield of hexane branched catalysts, the product distributions almost remain unchanged with time
isomers than Pt/H-MOR-p in the initial reaction stage. The available on stream (Fig. S6). Additionally, the selectivity for the branched iso-
Lewis acid sites due to the formation of extra-framework Al species in mers for all the investigated catalysts follows the order: 2, 2-di-
Pt/H-MOR-p-a have a promoting effect on the rates of the hydro- methylbutane < 2, 3-dimethylbutane < 3-methylpentane < 2-methyl-
isomerization, cracking and coking that leads to catalyst deactivation pentane, similar to the order of the molecular diameters of these hexane
[28]. Therefore, although the initial conversion of n-hexane is enhanced isomers. Moreover, Pt/H-MOR-p-a-a-a provides almost the same se-
over Pt/H-MOR-p-a in comparison with that over Pt/H-MOR-p, the lectivity for the branched isomers as Pt/H-MOR-p, indicating that the
catalytic activity will be rapidly dropped due to coke formation on the introduced mesopores do not alter the shape selectivity of Pt/H-MOR
catalysts. This observation agrees well with that most recently reported
[27].

4. Conclusions

The simple acid post-treatment applied over a commercial MOR


zeolite does not obviously introduce mesoporosity in the zeolite crystals
but it can significantly increase the Si/Al ratio due to partial deal-
umination that results in the formation of extra-framework Al species.
The prepared corresponding bifunctional Pt/H-MOR catalyst does not
show an improved catalytic activity in the hydroisomerization reaction
of n-hexane since the extra-framework Al species can enhance the for-
mation of cracking products and coke. These results may depend on the
specific MOR under investigation because other synthesis/post-treat-
ment procedures or commercial MOR may lead to different results re-
garding the acid leaching treatment. On the other hand, the sequential
acid-alkaline-acid post-treatment can introduce significant meso-
porosity in the MOR crystal structure. The prepared corresponding bi-
Fig. 2. Comparative results on the hydroisomerization of n-hexane over Pt/H- functional zeolite-supported Pt catalyst shows then an enhanced cata-
MOR-p (a), Pt/H-MOR-p-a (b) and Pt/H-MOR-p-a-a-a (c). Reaction conditions: lytic activity with long-term stability, due to the facts that the available
WHSV (n-hexane) = 2 h−1, H2/n-hexane molar ratio = 6:1, and T = 513 K. mesopores in the zeolite catalyst can shorten the diffusion length of the

24
X. Lu, et al. Catalysis Communications 125 (2019) 21–25

reactants and produced branched hexanes inside the micropores, and [9] L.Q. Meng, B. Mezari, M.G. Goesten, W. Wannapakdee, R. Pestman, L. Gao,
enhance the dispersion of Pt in the zeolite. The latter contributes to the J. Wiesfeld, E.J.M. Hensen, Catal. Sci. Technol. 7 (2017) 4520–4533.
[10] H. Chen, X. Shi, F. Zhou, H.X. Ma, K. Qiao, X.Y. Lu, J. Fu, H. Huang, Catal.
enhancement of proximity of Pt and acid sites in the mesopores of Commun. 110 (2018) 102–105.
zeolite which is a very important aspect in the present catalytic reac- [11] L.Z. Huang, P.F. Wang, J.F. Li, J.G. Wang, W.B. Fan, Microporous Mesoporous
tion. Mater. 223 (2016) 230–240.
[12] Z.X. Qin, B.J. Shen, X.H. Gao, F. Lin, B.J. Wang, C.M. Xu, J. Catal. 278 (2011)
266–275.
Acknowledgements [13] Z.X. Qin, K.A. Cychosz, G. Melinte, H.E. Siblani, J. Gilson, M. Thommes,
C. Fernandez, S. Mintova, O. Ersen, V. Valtchev, J. Am. Chem. Soc. 139 (2017)
17273–17276.
The authors gratefully acknowledge the financial support from the [14] K. Tarach, K. Góra-Marek, J. Tekla, K. Brylewska, J. Datka, K. Mlekodaj,
National Natural Science Foundation of China (20971109, 21036006, W. Makowski, M.C. Igualada López, J. Martínez Triguero, F. Rey, J. Catal. 312
21476214, and 21706239) and the Programme of Introducing Talents (2014) 46–57.
[15] C.R. Patil, P.S. Niphadkar, V.V. Bokade, P.N. Joshi, Catal. Commun. 43 (2014)
of Discipline to Universities (D17008).
188–191.
[16] J.C. Groen, W.D. Zhu, S. Brouwer, S.J. Huynink, F. Kapteijn, J.A. Moulijn, J. Pérez-
Appendix A. Supplementary data Ramírez, J. Am. Chem. Soc. 129 (2007) 355–360.
[17] S. Fathi, M. Sohrabi, C. Falamaki, Fuel 116 (2014) 529–537.
[18] H. Liu, S.J. Xie, W.J. Xin, S.L. Liu, L.Y. Xu, Catal. Sci. Technol. 6 (2016) 1328–1342.
Supplementary data to this article can be found online at https:// [19] D.R. Wang, L. Zhang, L. Chen, H.H. Wu, P. Wu, J. Mater. Chem. A 3 (2015)
doi.org/10.1016/j.catcom.2019.03.017. 3511–3521.
[20] J.C. Groen, L.A.A. Peffer, J.A. Moulijn, J. Pérez-Ramírez, Microporous Mesoporous
Mater. 69 (2004) 29–34.
References [21] A. Klein, R. Palkovits, Catal. Commun. 91 (2017) 72–75.
[22] J.C. Groen, J.C. Jansen, J.A. Moulijn, J. Pérez-Ramírez, J. Phys. Chem. B 108 (2004)
[1] R.A. Asuquo, G. Eder-Mirth, K. Seshan, J.A.Z. Pieterse, J.A. Lercher, J. Catal. 168 13062–13065.
(1997) 292–300. [23] X.F. Li, R. Prins, J.A. van Bokhoven, J. Catal. 262 (2009) 257–265.
[2] Y. Hong, J.J. Fripiat, Microporous Mater. 4 (1995) 323–334. [24] A.N.C. van Laak, S.L. Sagala, J. Zečević, H. Friedrich, P.E. de Jongh, K.P. de Jong, J.
[3] Y.S. Tao, H. Kanoh, L. Abrams, K. Kaneko, Chem. Rev. 106 (2006) 896–910. Catal. 276 (2010) 170–180.
[4] N. Batalha, L. Pinard, C. Bouchy, E. Guillon, M. Guisnet, J. Catal. 307 (2013) [25] H. Xu, Y.T. Zhang, H.H. Wu, Y.M. Liu, X.H. Li, J.G. Jiang, M.Y. He, P. Wu, J. Catal.
122–131. 281 (2011) 263–272.
[5] J. Zečević, G. Vanbutsele, K.P. de Jong, J.A. Martens, Nature 528 (2015) 245–254. [26] J. Pastvova, D. Kaucky, J. Moravkova, J. Rathousky, S. Sklenak, M. Vorokhta,
[6] G.D. Lei, B.T. Carvill, W.M.H. Sachtler, Appl. Catal. A Gen. 142 (1996) 347–359. L. Brabec, R. Pilar, I. Jakubec, E. Tabor, P. Klein, P. Sazama, ACS Catal. 7 (2017)
[7] D. Lozano-Castello, W.D. Zhu, A. Linares-Solano, F. Kapteijn, J.A. Moulijn, 5781–5795.
Microporous Mesoporous Mater. 92 (2006) 145–153. [27] P. Sazama, J. Pastvova, D. Kaucky, J. Moravkova, J. Rathousky, I. Jakubec,
[8] M. Tromp, J.A. van Bokhoven, M.T. Garriga Oostenbrink, J.H. Bitter, K.P. de Jong, G. Sadovska, J. Catal. 364 (2018) 262–270.
D.C. Koningsberger, J. Catal. 190 (2000) 209–214. [28] Q.L. Wang, G. Giannetto, M. Guisnet, J. Catal. 130 (1991) 471–482.

25

You might also like