You are on page 1of 13

Fuel 237 (2019) 178–190

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Hydrocracking of vacuum gas oil over NiMo/zeolite-Al2O3: Influence of T


zeolite properties

P.P. Dika, , I.G. Danilovaa, I.S. Golubeva, M.O. Kazakova, K.A. Nadeinaa, S.V. Budukvaa,
V.Yu. Pereymaa, O.V. Klimova, I.P. Prosvirina, E.Yu. Gerasimova, T.O. Bokc, I.V. Dobryakovab,
E.E. Knyazevab,c, I.I. Ivanovab,c, A.S. Noskova
a
Boreskov Institute of Catalysis SB RAS, Pr. Lavrentieva 5, 630090 Novosibirsk, Russia
b
Department of Chemistry, Moscow State University, Lenin Hills 1/3, 119991 Moscow, Russia
c
A.V. Topchiev Institute of Petrochemical Synthesis RAS, Leninskiy Prospect 29, 119991 Moscow, Russia

G R A P H I C A L A B S T R A C T

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

Keywords: Hydrocracking of vacuum gas oil has been studied over NiMo/zeolite-Al2O3 catalysts. Three different zeolites
Hydrocracking have been used for catalysts preparation: zeolites Beta (BEA) and Y (FAU) having small crystal size and zeolite Y
Vacuum gas oil modified by recrystallization (RFAU). HRTEM, low-temperature N2 adsorption, FTIR of adsorbed CO and TPD-
Catalyst NH3 showed that zeolites had different crystal sizes, mesopore volume, strength and concentration of acid sites.
Zeolite beta
Sulfide active component particles have been revealed to be similar in all catalysts by HRTEM and XPS. NiMo/
Micro-mesoporous zeolite Y
BEA catalyst having zeolite with the smallest average particle size and the highest concentration of Brønsted acid
Recrystallization
sites (BAS) demonstrated the highest hydrocracking activity. Selectivity to middle distillates decreased in the
following order: NiMo/FAU > NiMo/RFAU > NiMo/BEA. This effect is accounted for by optimal zeolite
acidity and improved availability of the acid sites for bulky molecules of the heavy feedstock.

1. Introduction products (gasoline, kerosene and low-sulfur diesel fuel) in the presence
of H2 and heterogeneous catalysts. Kerosene and diesel fractions ob-
Hydrocracking is an oil refining process that converts petroleum tained by hydrocracking have high quality due to high smoke point and
fractions with high-boiling point, such as vacuum gas oil, deasphalted cetane number, respectively, and low sulfur content. Therefore, middle
oil, atmospheric residue, and other heavy streams, into more useful distillates are the most valuable hydrocracking products [1,2].


Corresponding author.
E-mail address: dik@catalysis.ru (P.P. Dik).

https://doi.org/10.1016/j.fuel.2018.10.012
Received 17 March 2018; Received in revised form 12 September 2018; Accepted 1 October 2018
Available online 05 October 2018
0016-2361/ © 2018 Published by Elsevier Ltd.
P.P. Dik et al. Fuel 237 (2019) 178–190

Hydrocracking catalysts are bifunctional catalysts, which have two distillates of NiW hydrocracking catalysts have been studied in the ar-
types of sites: hydrogenating and acidic [3]. Sulfides of transition me- ticle [42], where it was shown, as in the article [40], that a decrease of
tals (Mo, W), promoted by nickel [4,5] or cobalt [6–8], or noble metals, zeolite crystals size resulted in an increase of selectivity to middle
such as Pt [9,10], Pd [11,12] and their bimetallic compounds [13,14], distillates and hydrocracking activity.
are generally used as the hydrogenating components of hydrocracking In a lot of works [17,40,43–47] zeolite Beta is used for hydro-
catalysts. Hydrogenating activity of noble metals is higher than that of cracking catalysts preparation. A direct comparison have shown that
the transition metal sulfides, but NiW and NiMo catalysts are widely the selectivity to middle distillates of hydrocracking catalysts based on
used in industry, because they are cheaper and more resistant to sulfur zeolite Beta was lower than that of the catalysts based on zeolite Y,
compounds. except the article [40], where zeolite Beta with a crystal size of 10 nm
Zeolites [4,10] and amorphous silica-alumina (ASA) [9,15], as well was used. It seems essential to compare different approaches for pre-
as their composites [16,17], are generally used as cracking components paration of hydrocracking catalysts, which are more active and selec-
for preparation of hydrocracking catalysts. Selectivity of zeolite-con- tive to middle distillates, by increasing availability of acid sites in
taining catalysts to the middle distillates is lower than selectivity of the zeolites. In this study, the influence of zeolites properties on catalysts
catalysts based on ASA [18,19]. This can be explained by the fact that performance in hydrocracking of VGO was investigated. Zeolites Beta
the strength and concentration of acid sites in zeolites is higher than in and Y having small crystal sizes and zeolite Y modified by re-
amorphous silica-alumina. Besides, the size of the heavy feed molecules crystallization were used for preparation of hydrocracking catalysts.
is often comparable or exceeds the size of the zeolite micropores.
Therefore, zeolite-containing catalysts are more active in hydrocracking 2. Experimental
of C10-C20 hydrocarbons, which results in overcracking of middle dis-
tillates and excessive production of light fractions [20]. 2.1. Preparation
Low activity of ASA can be compensated by an increase in process
temperature, but it often causes rapid deactivation of catalysts and a 2.1.1. Preparation of zeolites
decrease in the quality of the products due to the increase of aromatic Reaction mixture for zeolite Y with small crystal size (denoted as
compounds content [19,21]. Thus, an increase in the average pore size FAU) synthesis was prepared by mixing sodium silicate, sodium alu-
of zeolites combined with appropriate acidity and textural properties is minate, sodium hydroxide and silica. Zeolite FAU with SiO2/
necessary to enhance activity of catalysts in hydrocracking of heavy Al2O3 = 25 was prepared by consecutive procedures of hydrothermal
feedstocks [22]. For zeolite-containing catalysts, the concentration and crystallization at 95 °C, deep ion exchange in NH4NO3 solution for re-
strength of acid sites have a significant influence on activity and se- sidual sodium content 0,4 wt%, vapor treatment and acidic leaching
lectivity. Ding et al. [23] have shown that activity in hydrocracking of with nitric acid.
hexadecane increases with increasing acidity of the zeolite. Nowadays, Micro-mesoporous zeolite Y denoted as RFAU was prepared by hy-
ultra-stable zeolites Y are widely used as acidic component of com- drothermal treatment of dealuminated zeolite Y (CBV-720, Zeolyst) in a
mercial hydrocracking catalysts [24]. The unsteamed zeolite Y has a mixture of ammonia and hexadecyltrimethylammonium bromide aqu-
very high concentration of Brønsted acid sites, which results in very low eous solutions, according to [48].
selectivity to middle distillates. It is well known that an increase in the Zeolite Beta with small crystal size denoted as BEA was synthesized
Si/Al ratio in zeolites leads to a decrease in the total number of acid from reaction mixture prepared from aerosil A300 (Degussa), sodium
sites, as well as to an increase in their thermal stability [25]. To in- aluminate, sodium hydroxide and tetraethylammonum hydroxide (35%
crease a Si/Al ratio and selectivity to middle distillates, three different water solution, Sigma-Aldrich) as template. After hydrothermal crys-
methods of zeolites dealumination are used: hydrothermal treatment, tallization at 135 °C, following preparation steps include washing,
chemical treatment, and combination of hydrothermal and chemical drying, calcination for template removing and ion exchange in NH4NO3
treatment [26,27]. solution.
Another way to enhance selectivity of zeolites to middle distillates is
to form additional mesopores. It is known that mesopores are formed 2.1.2. Preparation of the supports and catalysts
during dealumination [28–30] and desilication [31]. Other promising AlOOH boehmite (ISCZC, Russia) and zeolites were used for pre-
approach to create mesoporosity in zeolites is recrystallization of zeo- paration of supports. Three supports with zeolite content of 30 wt%
lites with surfactants [32–38]. Recrystallization significantly increases were prepared by mixing one of the zeolite RFAU, FAU or BEA with
the volume of mesopores and the accessibility of acid sites. The number AlOOH followed by peptization with HNO3 and extrusion by plunger
and strength of the acid sites decrease with increasing the degree of extruder. Final trilobe extrudates were dried at 120 °C and calcined at
recrystallization, while the accessibility of the sites is improved. It was 550 °C.
shown previously for USY that the high selectivity of NiMo hydro- Catalysts were prepared by impregnation of the corresponding
cracking catalysts to middle distillates was the result of additional in- supports with aqueous solution prepared from nickel carbonate, am-
tracrystalline mesoporosity formed by recrystallization of the parent monium heptamolybdate and citric acid with mass ratio of the com-
zeolite [39]. ponents 1:2.9:2.5 (pH of the solution was 4). Impregnated catalysts
Another way to increase availability of acid sites is to reduce zeo- were dried at 120 °C and calcined at 550 °C. Citric acid used for cata-
lites crystals size. Camblor et al. [40] have reported that activity and lysts preparation because: it allows stable active metals solution to be
selectivity to middle distillates of hydrocracking catalysts based on prepared; Ni and Mo are located in close proximity due to complex
zeolite Beta increases with decreasing size of zeolite crystals. Authors formation [49]; favors high dispersion of active metals due to high
showed that activity of the catalyst based on the zeolite Beta with a viscosity of solution on the drying step [50]. Thus, the use of citric acid
crystal size of 10 nm was significantly higher than that of the catalyst for catalysts preparation, even with the subsequent calcination of the
based on zeolite Y. Moreover, the selectivity to middle distillates of the catalysts, improves the deposition of active metals.
catalyst based on the nanocrystalline zeolite Beta was higher than the Ni and Mo content in the catalysts was 3.0 ± 0.1 and 9.8 ± 0.5 wt
selectivity of zeolite Y based catalyst and even higher than that of ASA- %, respectively. The catalysts were denoted as NiMo/FAU, NiMo/RFAU
based catalyst. Landau et al. [41] have found that activity of Pt con- and NiMo/BEA.
taining hydrocracking catalyst based on zeolite Beta with crystal sizes Also, two additional catalysts – NiMo/Al2O3 and NiW/ASA-Al2O3
of 10–30 nm is much higher than that of the hydrocracking catalyst were prepared. NiMo/Al2O3 catalyst and NiW/ASA-Al2O3 were used as
based on zeolite Beta with crystal sizes of 200–500 nm. The influence of the first and second layers in a stacked bed to simulate industrial
zeolite Y crystals size on the activity and selectivity to the middle conditions in laboratory unit. NiMo/Al2O3 was prepared using similar

179
P.P. Dik et al. Fuel 237 (2019) 178–190

impregnation solution, as described in [51]. NiMo/Al2O3 was used as the H-complexes with the CO molecules using the following coefficients
hydrotreating catalyst. Therefore, to enhance hydrotreating activity the (A0 = 54 cm/μmol for the complexes with νOHOH⋯CO 3190–3300 cm−1
catalyst was only dried at 120 °C instead of calcination at 550 °C. Ni and and A0 = 27 cm/μmol for complexes with νOHOH⋯CO 3385–3500 cm−1).
Mo content in NiMo/Al2O3 catalyst was 3.7 and 12.5 wt% respectively. Good correlation between the concentration of acid sites determined by
NiW/ASA-Al2O3 catalyst was prepared by impregnation with the aqu- ammonia and by CO was shown in ref. [54]. The shift of IR band of the
eous solution prepared from nickel carbonate, ammonium para- acidic hydroxyls groups (ΔνOHOH⋯CO) were recalculated into the value
tungstate and citric acid followed by subsequent drying at 120 °C and of proton affinity (PA): PA [kJ/mol] = 1390–442.5*log10(ΔνOHOH⋯CO/
calcination at 550 °C [52]. Ni and W content in NiW/ASA-Al2O3 catalyst (ΔνOHSiOH⋯CO), where 1390 kJ/mol corresponds to the PA of surface
was 3.1 and 17.4 wt% respectively. OH-groups of Aerosil [55]. The concentration of BAS (method-II) was
For XPS and HRTEM analysis the catalysts were sulfided as follows: evaluated from the integral intensity (A) of CO band in the range of
1 g of a catalyst was placed in the glass flow reactor and sulfided during 2170–2180 cm−1 using the equation C = A/A0 with the molar integral
2 h at 220 °C and 2 h at 400 °C in the 20 cm3/min H2S flow. absorption coefficients values A0 = 2.7 cm/μmol [56]. The concentra-
tion of Lewis acid sites (LAS) was evaluated from the integral intensity
2.2. Characterization of CO band in the range of 2180–2233 cm−1. The following molar in-
tegral absorption coefficients values (A0, cm/μmol) were used: 1.23
Elemental analysis of the catalysts was carried out using the method (2233–2225 cm−1), 1.1 (2216–2206 cm−1), 0.9 (2198–2190 cm−1)
of atomic emission spectroscopy with inductively coupled plasma (ICP- [57]. The concentration of OH groups was assessed from the integral
AES) on Optima 4300 DV. intensities of absorption bands in the hydroxyls stretching region using
Low temperature adsorption-desorption of N2 for zeolites was following coefficients (A0 = 7.5 and 4.7 cm/μmol for strongly acidic
measured at 77 K using an automated porosimeter (ASAP 2000 bridged hydroxyl groups in the faujasite supercages and sodalite cages,
Micromeritics). The BET method was applied to calculate the total correspondingly [56]. In the presented spectra, the absorbance was
surface area, which was used for comparative purposes. The t-plot normalized to sample wafer density.
method was used for determination of micropore volume. The total X-ray photoelectron spectra (XPS) were recorded using a SPECS
sorbed volumes, including adsorption in the micropores and mesopores, spectrometer (Germany) with a PHOIBOS-150 hemispherical energy
were calculated from the amount of nitrogen adsorbed at a relative analyser and AlKα irradiation (hν = 1486.6 eV, 200 W). The binding
pressure P/P0 of 0.95, before the onset of interparticle condensation. energy scale was preliminarily calibrated using the peak positions of the
Pore size distribution was calculated by BJH method using the ad- Au4f7/2 (84.0 eV) and Cu2p3/2 (932.67 eV) core levels. The samples
sorption branch of the isotherm. were supported using conductive scotch tape. The internal reference
Textural properties of the catalysts were determined by nitrogen method was used for the correct calibration of the photoelectron peaks.
physisorption using an ASAP 2400 (USA) instrument. BET surface areas The C1s peak (Eb = 284.8 eV) corresponded to surface hydrocarbon-
were calculated from the amount of nitrogen uptaken at relative pres- like deposits (C–C and C–H bonds) accumulated on the surface during
sures that ranged from 0.05 to 0.30. The total pore volume was derived the storage in the atmosphere. A low-energy electron gun (FG-15/40,
from the amount of nitrogen adsorbed at a relative pressure close to SPECS) was used for the sample charge neutralization.
unity (in practice, P/P0 = 0.995) under the assumption that all of the
accessible pores had been filled with condensed nitrogen in the normal 2.3. Catalytic testing and product analysis
liquid state. The t-plot method was used for determination of micropore
volume. Testing in hydrocracking of vacuum gas oil was carried out in a
HRTEM images were obtained on a JEM-2010 electron microscope high-pressure trickle-bed unit. Catalysts in the form of trilobe granules
(JEOL) with a lattice-fringe resolution of 0.14 nm and operated at an with a length of 3–6 mm and without defects were used for the loading
accelerating voltage of 200 kV. The high-resolution images of periodic in the reactor. The total volume of catalysts in the reactor was 60 cm3:
structures were analyzed by the Fourier method. Samples to be ex- the first layer – NiMo/Al2O3 20 cm3; the second layer – NiW/ASA-Al2O3
amined by HRTEM were prepared on a perforated carbon film mounted 20 cm3; the third layer – NiMo/zeolite 20 cm3. Catalyst granules were
on a copper grid. diluted with fine sized SiC (0.15–0.3 mm). This approach is widely used
Temperature programmed desorption of ammonia (TPD-NH3) was for testing of granulated hydrotreating and hydrocracking catalysts.
performed on chemisorption analyzer USGA-101 (production of UNISIT Dilution of granules with inert fine powders improves the wetting of
Ltd., Russia) equipped with TC detector. The NH3 adsorption was car- granules, reduce the wall effect, etc. [58]. A mixed feed comprising
ried out during 30 min, at ambient temperature in a flow of NH3 diluted straight-run VGO (69 wt%), heavy coker gas oil (22 wt%) and fractions
with N2 (1/1). Subsequently, the physically adsorbed NH3 was removed from solvent extraction unit (aromatic extract – 7 wt%) and solvent
in a flow of dry He at 373 K for 1 h. Typical TPD experiments were dewaxing unit (petrolatum – 2 wt%), was used as a feed. The properties
carried out in the temperature range of 293–1053 K in a flow of dry He of the feedstock used are shown in Table 1. Catalysts were sulfided in
(30 cm3/min). The rate of heating was 8 K/min. situ at a pressure of 3.5 MPa using the mixture of straight run diesel
Surface acid sites of the zeolites were also characterized by IR
spectroscopy of adsorbed carbon monooxide. IR spectra were recorded Table 1
on a Shimadzu FTIR-8300 spectrometer within the spectral range of Properties of feedstock.
700–6000 cm−1 with a resolution of 4 cm−1 and 500 scans for signal Characteristic Value
accumulation. The powder samples were pressed into thin self-sup-
porting wafers (0.004–0.006 g/cm2) and activated in the special IR cell a
Boiling range , °C
at 773 K for 2 h in dynamic vacuum of 10−3 mbar. CO was introduced 5–10 wt% 306–342
20–30 wt% 393–434
at liquid nitrogen temperature by doses up to an equilibrium pressure of 40–50 wt% 460–476
13 mbar. The strength of Brønsted acid sites (BAS) was estimated by the 60–70 wt% 489–502
method of hydrogen bonds based on the change in the stretching vi- 80–90 wt% 515–533
bration frequency of the OH groups that occurred under CO absorption. 95 wt% 549
Sulfur content, wt% 0.966
The higher the shift of IR band of OH stretching vibration of the hy-
Nitrogen content, wt% 0.109
droxyls groups (ΔνOHOH⋯CO), the stronger is the acidity of OH group Density (20 °C), g/cm3 0.927
[53]. The concentration of BAS (method-I) was determined from the
a
integral intensity of the band attributed to corresponding OH–group in According to ASTM D7213.

180
P.P. Dik et al. Fuel 237 (2019) 178–190

fraction, dimethyl disulfide and aniline. Aniline or other amines are


used for reversible passivation of the zeolite acid sites. This approach
protects the acid sites from the coking during the sulfidation stage when
active sulfide sites are not yet formed. This technique is described
previously [27,59,60]. The catalyst loading and sulfidation procedures
are thoroughly described in [15]. Hydrocracking tests were carried out
at a pressure of 16.0 MPa, a liquid hourly space velocity (LHSV) of
0.71 h−1 and H2 to oil ratio of 1500 (v/v). The temperature in the re-
actor was 360 °C during the first 24 h, then 380 °C during next 48 h and
then the temperature was increased up to 390 °C and higher. Each ex-
perimental temperature 390 and 410 °C (405 °C for NiMo/BEA) was
maintained for 192 h. Gas effluent from the separator was analyzed by
gas chromatography using FID and capillary column. Liquid products
were analyzed by SIMDIS-GC in accordance with the ASTM D7213
standard test method. Product yields were defined by summing up an
amount of fractions, determined by SIMDIS-GC, and an amount of
fractions from the gas effluent from the separator. The conversion of
Fig. 2. Pore size distribution for the zeolites.
VGO (XVGO), selectivity to gas (SGas ), selectivity to naphtha (Snaphtha )
and selectivity to middle distillates (SMD) were calculated using the
following equations: which is typical for zeolites with micro-mesoporous structure. FAU
zeolite has H3 type of the loop. Finally, it is obvious that zeolites are
XVGO (%) = 100%−360 °C+ in product (wt%)
very different in the pore structure, however, it can be suggested that
C1−C4 in product (wt%) all zeolites contain micro and mesopores. Pore size distribution of the
SGas (%) = × 100% zeolites is shown in Fig. 2.
100%−360 °C+ in product (wt%)
The specific surface area of the zeolites determined by the BET
C5−130 °C in product (wt%) method and the micropore volume of the zeolites determined by t-plot
SNaphtha (%) = × 100%
100%−360 °C+ in product (wt%) method are given in Table 2. The highest mesopore volume is observed
for the RFAU zeolite obtained by recrystallization. According to XRD
130−360 °C in product (wt%) data published earlier [39] crystal structure of zeolite RFAU is pre-
SMD (%) = × 100%
100%−360 °C+ in product (wt%) served during recrystallization, whereas the crystallinity is slightly re-
duced. The FAU zeolite has higher mesopore volume than the BEA
zeolite.
3. Results and discussion
The crystals size of the zeolites was determined from HRTEM
images, at least 200 particles for each sample were used to calculate the
3.1. Chemical composition and textural properties of zeolite materials and
average size of the zeolite crystals and to obtain the particle size dis-
catalysts
tribution (Fig. 3). According to HRTEM data, the surface of zeolite
crystals of RFAU is observed to be covered by a film of worm-like
The zeolites and the catalysts in oxide form were investigated by
mesoporous solid. Zeolite FAU has well-crystallized particles with mi-
low-temperature nitrogen adsorption. Nitrogen adsorption-desorption
croporous structure as well as mesopores formed by dealumination, but
isotherms of the zeolites are shown in Fig. 1. The zeolites have type IV
this sample has much wider range of crystal sizes than RFAU sample, as
adsorption/desorption isotherms with a distinct hysteresis loops. The
it can be seen in Fig. 3. Unlike to RFAU zeolite, FAU zeolite has about
presence of the bend at low pressures is due to micropores in zeolites.
20% of crystals with the sizes lower than 200 nm. BEA zeolite has the
Difference in the height of the bend and whole isotherms indicates the
smallest crystal size – more than 85% of the crystals have the sizes
difference in the amount of pore with definite diameters. All isotherms
lower than 200 nm.
contain hysteresis loops caused by mesoporosity of the samples. RFAU
Catalysts in the oxide state were analyzed by atomic emission
zeolite show H1 type of the hysteresis loop, which exhibits a narrow
spectroscopy with inductively coupled plasma (ICP-AES). It was shown
range of uniform mesopores. BEA zeolite show H4 hysteresis loop,
that all catalysts have similar content of Ni and Mo (Table 3). Specific
surface area and pore volume of the catalysts do not correlate with the
specific surface and pore volume of the zeolites used for catalysts pre-
paration, as it follows from Tables 2 and 3.
Specific surface area of the initial zeolites increases in the following
order: BEA, RFAU, FAU, but specific surface area of the catalysts in-
creases in the following order: NiMo/BEA, NiMo/FAU, NiMo/RFAU.
Total pore volume of zeolites increases in the following order: BEA,
FAU, RFAU, but total pore volume of the catalysts increases in the

Table 2
Chemical composition and textural characteristics of the zeolites.
Zeolite Mole Vmicro, Vmeso, Vtotal, SBET, Average size of
ratio cm3/g cm3/g cm3/g m2/g the crystals,
SiO2/ nm
Al2O3

FAU 25 0.27 0.21 0.48 800 390


RFAU 28 0.17 0.39 0.56 760 550
BEA 29 0.22 0.15 0.37 660 160
Fig. 1. Nitrogen adsorption-desorption isotherms for the zeolites.

181
P.P. Dik et al. Fuel 237 (2019) 178–190

Fig. 3. Crystal size distribution for the zeolites.

Table 3
Chemical composition and textural characteristics of the catalysts.
Catalyst Content, wt.% Textural characteristics Fig. 4. IR spectra of zeolite samples in the OH stretching region: 1 – BEA; 2
–FAU; 3 – RFAU.
Ni Mo SBET, m2/g Vtotal, cm3/g Vmicro, cm3/g

NiMo/FAU 2.9 9.7 259 0.39 0.052


Table 4
NiMo/RFAU 2.9 9.8 298 0.42 0.049
Type and concentration of some OH-groups of zeolites.
NiMo/BEA 3.1 10.1 250 0.44 0.056
Type of sites Absorbance Zeolite samples
bands, cm−1
following order: NiMo/FAU, NiMo/RFAU, NiMo/BEA. BEA RFAU FAU
Probably, despite the same method used for the preparation of the
Integrated intensity of IR bands
supports and the catalysts, pores blocking of the different zeolites with corresponded to some OH-groups, cm−1
alumina and active-metals oxides occurs differently.
Si–OH terminal νOH 3746 0.720 1.525 1.075
Si–OH internal or defect νOH 3738 0.880 0.625 0.835
3.2. Nature and concentration of acid sites of zeolites Si–O(H)–Al νOH 3550–3630 0.565* HF – HF –
0.470** 0.435**
3.2.1. IR CO total – total –
1.265*** 1.545***
Fig. 4 presents the IR spectra in the O–H stretching region of three
Al–OH extraframework νOH 3670–3720 0.560 0.830 1.050
zeolites under investigation. The IR spectra of BEA zeolite (Fig. 4, sp. 1) Al–OH extraframework, νOH 3775–3790 0.115 0.030 0.015
has one band at 3610 cm−1 that can be assigned to acidic bridged Si–O terminal
(H)–Al hydroxyl groups. The wide band at 3300–3550 cm−1 refers to
hydroxyl groups perturbed by a hydrogen bond. The position of the IR
*
– Si–O(H)–Al groups, νOH 3610 cm−1.
bands of bridged hydroxyls is in good agreement with the literature
**
– Si–O(H)–Al groups in supercages, HF, νOH 3630 cm−1.
***
– total Si–O(H)–Al groups, νOH 3550–3630 cm−1.
data for BEA zeolites [61–63]. The spectra of FAU and RFAU samples
(Fig. 4, sp. 2 and 3) show the two intense bands with the maxima at ca.
attributed to low-acidic hydroxyl groups bonded to extra-lattice alu-
3630 and 3570 cm−1 that are typical for IR spectra of USY zeolites
minium [61,62,67].
[56,64–66], and assigned to the stretching vibrations of the two main
Varying the amount of different types of OH-groups in the in-
kinds of «structural» hydroxyl groups – bridged Si–O(H)–Al groups in
vestigated samples is shown in the Table 4. The nature of the OH-
the supercages (HF, high frequency) and bridged OH in sodalite cages
groups in micro-mesoporous RFAU and zeolite FAU with small crystal
(LF, low frequency), respectively. The less intense bands at 3600 and
size samples is close, but the number of some hydroxyls is dissimilar.
shoulder at 3550 cm−1 are mostly attributed to HF and LF groups of
The FAU zeolite is characterized by a lower concentration of acidic
faujasites polarized by Lewis acidic extraframework Al species [64,66].
bridged Si–O(H)–Al hydroxyl groups in the supercages and terminal
The signal at 3746 cm−1 in the spectra of all studied zeolite samples
silanols, but by a higher concentration of defect Si–OH and acidic ex-
refers to terminal silanol OH groups while shoulder at 3738 cm−1
traframework Al–OH groups than RFAU sample. The BEA zeolite
corresponds to internal or defect silanol OH groups typical for zeolites
sample has the maximum concentration of low-acidic terminal Al-OH
and MCM-41 aluminosilicates, which are located in the close vicinity to
groups bonded to extraframework aluminium. The number of acidic
the lattice imperfection or Lewis acid site (e.g., tricoordinated Al atom,
bridged Si–O(H)–Al hydroxyl groups on the BEA surface is higher than
Si–OH⋯Al3+ groups) [67,68]. Previously we proposed that these defect
the amount of the same groups in the faujasite supercages. It can be
silanol groups can be located both at the external surface of zeolite
assumed that bridged hydroxyl groups in other faujasite cages are in-
crystallites and in the intracrystalline mesopores or in the channels of
accessible to bulky molecules of heavy feedstock.
MCM-41-like material of RFAU [38]. Other less intense bands at 3670,
CO is often used as a suitable probe molecule to study Brønsted
3695 and 3720 cm−1 can possibly be assigned to the acid OH groups
acidity of zeolites. Due to H-bonding with Brønsted acid sites, CO in-
attached to tricoordinated Al atoms, which are partially connected to
duces a broadening and a red shift of the OH bands. Simultaneously,
the framework, while signals and 3775 and 3790 cm−1 can be

182
P.P. Dik et al. Fuel 237 (2019) 178–190

USY zeolites [52,57] and for materials obtained by zeolite Y re-


crystallization [38], and to the shift at ΔνO–H⋯CO = −380 cm−1 for LF
band [57] located possibly in destroyed sodalite cages. At the same
time, stretching frequency of CO bond undergoes blue shift to
∼2180–2182 cm−1 (ΔνCO = +37–39 cm−1) (Fig. 6-A and B). It should
be noted that the strength of BAS is similar for FAU and RFAU zeolite
samples. The appearance of small negative peaks at ca. 3550 and
3600 cm−1 in difference spectra after CO adsorption can be due to
perturbation of a few bridging Si–O(H)–Al groups polarized by LAS
species. The band at 3738 cm−1, which is observed as a shoulder to the
band at 3746 cm−1 in initial spectrum, exhibits also a large shift af-
fording a positive peak at ca. 3475 cm−1 (ΔνO–H⋯CO = −265 cm−1). It
demonstrates that corresponding Si–O(H)⋯Al3+ defect groups are less
acidic than bridging Si–O(H)–Al groups in faujasite cages according to
the value of the red shift of OH vibrations with adsorbed CO. Corre-
sponding CO band is at ∼2170 cm−1 (ΔνCO = +27 cm−1). The ap-
pearance of positive bands at 3420* cm−1 can be due to perturbation of
the acid OH groups attached to tricoordinated Al atoms, which are
partially connected to the framework (perhaps bands at ca.
3670–3695 cm−1) [60]. Besides, the highest pressure of CO close to
equilibrium results in the shift in the intensity of the silanol bands from
3746 cm−1 to ca. 3660 cm−1: a shift of ∼−90 cm−1 is typical for weak
acidic terminal Si–OH groups [52]. Corresponding CO band is at
∼2157 cm−1 (ΔνCO = +14 cm−1).
Fig. 5. IR difference spectra in the OH-region between the initial zeolites BEA CO adsorption on BEA sample gives rise to the shift of the band at
(1); FAU (2); RFAU (3) and those with 3 mbar of adsorbed CO at liquid nitrogen 3610 cm−1 (bridged Si–O(H)–Al hydroxyl groups in the BEA channels)
temperature (1 mbar for spectrum 1a). to 3270 cm−1 at the low CO pressure (Fig. 5, sp. 1a) and to the shift of
the band at 3625 cm−1 (bridged Si–O(H)–Al hydroxyl groups in a larger
channels) to 3290 cm−1 at the higher CO pressure (Fig. 5, sp. 1). The
C^O stretching frequency shows the blue shift relative to the gas phase
magnitude of the red shift is about −340 cm−1, which is close to the
CO (νCO = 2143 cm−1). The higher the OH acidity, the larger the shift
observed strength of the bridged groups of conventional and re-
of the OH modes and the higher the carbonyl stretching frequency.
crystallized H-Beta zeolites [34,60]. At the same time CO stretching
Progressive CO adsorption was carried out for all zeolites (Figs. 5
frequency undergoes blue shift to ∼2179 cm−1 (ΔνCO = +36 cm−1)
and 6). During low temperature CO adsorption on FAU and RFAU
(Fig. 5-C). It can be noted that the strength of structural bridging groups
samples, the HF band fully disappeared, and the new band at
for BEA zeolite (ΔνO–H⋯CO = −340 cm−1) is insignificantly lower than
3280 cm−1 appeared, but LF band was only partially perturbed and
for FAU and RFAU samples (ΔνO−H⋯CO = −350 cm−1). The appear-
shifted down to 3190 cm−1 (Fig. 5, sp. 2 and 3). This corresponds to the
ance of positive peaks at ca. 3450 cm−1 in difference spectra after CO
shift at ΔνO–H⋯CO = −350 cm−1 for HF band, that is typical value for
adsorption can be due to perturbation of Si–O(H)⋯Al3+ defect groups

Fig. 6. IR spectra of CO adsorbed at liquid nitrogen temperature on A – RFAU; B – FAU; C – BEA zeolites with increased dosage of adsorbed CO: PCO – 0.1, 0.2, 0.3,
0.4, 0.6, 1, 2, 4, 6 mbar.

183
P.P. Dik et al. Fuel 237 (2019) 178–190

Table 5
Type and concentration of some acid sites of zeolites by FTIR of adsorbed CO.
Type of sites Absorbance bands, cm−1 Acid sites strength Zeolite samples

BEA RFAU FAU

BAS concentration, μmol/g


Strong BAS in faujasite cages, νO–H = 3570–3630 cm−1 νO–H⋯CO = 3190–3280 PA = 1140–1125 kJ/mol, – 38a (65)b 40a (66)b
(νCO=2180) ΔνO–H⋯CO = −350–380 cm−1
Strong BAS in BEA channels, νO–H = 3610 cm−1 νO–H⋯CO = 3270–3290 PA = 1145–1150 kJ/mol, 85a – –
ΔνO–H⋯CO = −340–335 cm−1
BAS at external surface of microcrystallites or in νO–H⋯CO = 3450–3475 PA = 1180–1200 kJ/mol 37 a
25 a
40a
mesopores, νO–H = 3738 cm−1 ΔνO–H⋯CO = −280–260 cm−1
Total BAS νCO = 2180–2170 250b 125b 138b

LAS concentration, μmol/g


Strong LAS νCO = 2225–2235 QCO = 56.5–51.5 kJ/mol 53 33 27
Total LAS νCO = 2189–2235 173 84 57

Total acid sites concentration, μmol/g


Total acid sites (BAS + LAS) 423 209 195

a
– Concentration of BAS was evaluated by method I from the integral intensity of band attributed to corresponding OH–group in the H-complexes with the CO
molecules CO band (νO–H⋯CO).
b
– Concentration of BAS was evaluated by method II from the integral intensity of CO band (νCO).

(ΔνO–H⋯CO = −280 cm−1). Corresponding CO band is at ∼2175 cm−1


(ΔνCO = +32 cm−1). The terminal silanol groups when interacting
with CO show pertubation band at ∼3660 cm−1
(ΔνO–H⋯CO = −90 cm−1). Corresponding CO band is at ∼2157 cm−1
(ΔνCO = +14 cm−1).
According to IR spectra of adsorbed CO in carbonyl region, surface
groups in all the samples can be attributed to three types of LAS and
two types of BAS. The adsorption of CO (Fig. 6) on all the samples leads
to the appearance of the following absorbance bands: (1)
2235–2225 cm−1 relating to CO complexes with Al3+ ions in penta-
hedron environment being specific structure defects of zeolites (strong
LAS), (2) 2216–2210 cm−1 corresponding to LAS of medium strength,
(3) 2198–2192 cm−1 relating to CO complex with Al3+ ions of alumina
clusters (weak LAS) [53], (4) 2175, 2179–2182 cm−1 corresponding to
CO complex with strong BAS, (5) 2170 cm−1 relating to CO complex
with medium BAS (due to the content of Na2O < 0.02 wt%) and (6)
2157 cm−1 corresponding to CO complex with Si–OH groups [52].
The concentration of BAS and LAS for the zeolite samples is given in
Table 5. It can be seen that the concentration of strong BAS (PA
Fig. 7. Ammonia temperature-programmed desorption profiles for the zeolites.
1150–1125 kJ/mol) in faujasite cages or BEA channels (bridged Si–O
(H)–Al groups) decreases in the following order: BEA > FAU ≥ RFAU,
but the concentration of less acidic BAS (PA 1180–1200 kJ/mol) at the peak in the region of 100–350 °C due to ammonia desorption from weak
crystallites external surface and/or in the intracrystalline mesopores acid sites and the high-temperature peak in the 350–700 °C due to
(Si–O(H)⋯Al3+ defect groups) increases in the following order: ammonia desorption from strong acid sites.
RFAU < BEA ≤ FAU (concentration of BAS was evaluated by method The BEA sample has the highest concentration of strong and weak
I). Total BAS concentration decreases in the following order: BEA > acid sites (Table 6). Strength and concentration of acid sites in FAU and
FAU ≥ RFAU (concentration of BAS was evaluated by method II). It RFAU zeolites according to TPD-NH3 are approximately the same. The
can be seen that both the concentrations of strong LAS maximum of low-temperature peak for zeolite BEA has a higher tem-
(QCO = 56.5–51.5 kJ/mol) and total LAS decrease in the following perature, which may indicate that the weak acid sites of this zeolite are
order: BEA > RFAU > FAU (concentration of LAS was evaluated by stronger than of FAU and RFAU ones.
method II). The total Brønsted acid site density of the sample decreases The maxima of high-temperature peaks for all zeolites are ap-
in the following order: BEA (0.38 μmol/m2) > FAU (0.17 μmol/ proximately the same, which may indicate that strong acid sites in the
m2) ≥ RFAU (0.16 μmol/m2). The total acid site density of the sample zeolites have similar strength.
decreases in the following order: BEA (0.64 μmol/m2) > RFAU
(0.27 μmol/m2) > FAU (0.24 μmol/m2). Table 6
Concentration of acid sites according to TPD-NH3.

3.2.2. TPD-NH3 Zeolite Tmax °C Concentration of acid sites, µmol/g Total


acid site
The zeolite samples were investigated by TPD-NH3 method. The Weak Strong Weak Strong Total density,
thermodesorption curves of ammonia are shown in Fig. 7. (100–350 °C) (350–700 °C) (100–700 °C) µmol/
Brønsted and Lewis acid sites cannot be distinguished by TPD-NH3 m2
method, but TPD-NH3 allows a qualitative evaluation of the acid sites
FAU 185 372 280 220 500 0.63
strength: the higher the desorption temperature of ammonia, the
RFAU 182 362 320 190 510 0.67
stronger the acid sites. According to the TPD-NH3, all the investigated BEA 206 360 660 340 1000 1.52
zeolites have two forms of ammonia desorption: the low-temperature

184
P.P. Dik et al. Fuel 237 (2019) 178–190

Fig. 8. HRTEM images of the sulfide catalysts.

3.3. HRTEM of the catalysts 2.1–2.4 nm (Fig. 9a) and average stacking number of 2.0–2.1 (Fig. 9b)
are typical for sulfided supported catalysts described in the literature,
The catalysts in the sulfide form were investigated by HRTEM which contain the same active metal concentration as in our case.
method. Fig. 8 shows microphotographs of the catalysts. Zeolites
crystals in all the catalysts are decorated with γ-Al2O3 particles as it is
3.4. XPS
seen in the images with 1 μm scale. At the higher resolution, highly
dispersed MoS2 particles are visualized over the γ-Al2O3 surface. No
Analysis of the S 2s + Mo 3d XPS spectra of the studied samples
visible MoS2 particles were found over zeolites surface.
(Fig. 10, left) shows that the peak with the lowest binding energy
Additionally a small amount of bulk particles of active metal sul-
(BE = 226.4 eV) belongs to sulfur (the S 2s level). In order to determine
fides have been found in the HRTEM images for all the catalysts. The
the ratio between the sulfide and oxide forms of molybdenum more
quantitative estimation of the number of bulk particles is difficult due
accurately, we performed the decomposition of the Mo 3d line into
to their small amount. However, the NiMo/RFAU catalyst contains
individual spectral components with the use of the XPS Peak 4.1 soft-
somewhat more bulk particles than the NiMo/FAU and NiMo/BEA
ware. The contribution of S 2s peaks to the total S 2s + Mo 3d spectrum
catalysts.
was taken into account. Binding energies and integral intensities were
The sulfided active component was shown to be in the form of the
determined after subtraction of a Shirley-type background from initial
particles with similar morphology in the sulfided NiMo/FAU, NiMo/
photoelectron spectra. The Mo 3d spectra exhibited three main peaks
RFAU, and NiMo/BEA catalysts (Fig. 9). The average slab length of
which can be attributed to molybdenum in the sulfide environment

Fig. 9. Morphology of MoS2 particles: a) Slab length distribution b) Stacking number distribution. L – average slab length, N – average stacking number.

185
P.P. Dik et al. Fuel 237 (2019) 178–190

Fig. 10. S 2s + Mo 3d (left) and Ni 2p (right) XPS spectra of the sulfide catalysts.

(MoS2 species) characterized by BE (Mo4+) = 228.8 eV, Mo oxysulfide with HRTEM data, i.e. sample NiMo/RFAU has somewhat more bulk
with BE (Mo5+) = 230.5 ± 0.1 eV and molybdenum in the state of sulfide particles compared with that in NiMo/FAU and NiMo/BEA.
Mo6+ in the oxygen environment, for example, MoOx with BE
(Mo6+) = 232.6 ± 0.1 eV [69–73]. Integration of corresponding peaks 3.5. Hydrocracking activity and selectivity
shows that the sulfided catalysts contain 81–84% of Mo in the form of
MoS2, 13–14% in the form of Mo5+ and 3–5% in the form of Mo6+ Catalysts were sulfided and evaluated for catalytic performance in
(Fig. 10, left). The results of Ni 2p XPS spectra decomposition for the VGO hydrocracking. Catalytic tests were carried out with non-pre-
sulfided catalysts are shown in Fig. 10 (right). According to the litera- treated feedstock (Table 1) and the studied catalysts were loaded in
ture data [74,75], three components can be attributed to Ni in NiS stacked beds containing NiMo/Al2O3 as the top layer, NiW/ASA-Al2O3
(BE = 852.7 ± 0.1 eV), Ni in NiMoS phase (BE = 853.9 ± 0.1 eV) as the interlayer and NiMo/zeolite-Al2O3 hydrocracking catalyst as the
and Ni2+ species (BE = 856.2 ± 0.1 eV). It was found that the sulfide bottom layer. The top layer and the interlayer catalysts were used to
catalysts contain 61–69% of nickel in the form of NiMoS phase, 13–17% remove sulfur and nitrogen containing compounds in order to prevent
in the form of NiS and 18–22% in the form of Ni2+ (Fig. 10, right). poisoning of zeolite catalysts. In this case, hydrocracking catalysts were
NiMo/BEA and NiMo/FAU samples are characterized by close con- tested in the presence of small amounts of unconverted organic het-
centrations of any of nickel species, while NiMo/RFAU sample reveals erocompounds and NH3, H2S, some light hydrocarbons, which were
lower relative content of active NiMoS phase and higher content of Ni produced during hydrotreating. Such tests simulate once-through hy-
in the low active states (NiS and Ni2+). This observation agrees well drocracking process.

186
P.P. Dik et al. Fuel 237 (2019) 178–190

Table 7 RFAU zeolite should also be high due to mesopores formation by re-
The results of VGO hydrocracking: VGO conversion, selectivities to gas, to crystallization procedure. Mesopore volume in the RFAU zeolite is more
naphtha, to middle distillates and products yield. than 1.5 times higher than in the BEA and FAU zeolites.
Catalyst NiMo/FAU NiMo/RFAU NiMo/BEA In our previous work [39] it was shown that the catalyst prepared
with recrystallized zeolite Y had greater activity and selectivity to
Temperature, °C 390 410 390 410 390 405 middle distillates than the one prepared with parent non-crystallized
VGO conversion, % 34.4 89.1 30.6 73.0 35.6 97.8 zeolite Y. In hydrocracking process, catalyst selectivity is a measure of
Products yield, wt.% the wt. % (or vol. %) of feedstock converted to desired product or
Gas C1-C4 0.6 5.0 0.4 4.0 0.8 15.4 products. Product selectivities obtained with different catalysts are
Naphtha < 130 °C 2.5 26.3 1.6 19.1 5.3 45.0 often compared at the same conversion [76]. Therefore, to compare the
MD* 130–360 °C 31.3 57.8 28.5 49.9 29.5 37.4
catalysts selectivities, it is suitable to use the dependence of selectivity
UCO** > 360 °C 65.6 10.9 69.4 27.0 64.4 2.2
on VGO conversion. The most valuable products of hydrocracking are
Products selectivity, % middle distillates. Increase of VGO conversion leads to sharp decline of
Gas selectivity 1.7 5.6 1.3 5.5 2.2 15.7
Naphtha selectivity 7.3 29.5 5.2 26.2 14.9 46.0
selectivity to middle distillates for all catalysts (Fig. 11d). Selectivity to
MD selectivity 91.0 64.9 93.1 68.4 82.9 38.2 middle distillates increases in the following order of the catalysts:
NiMo/BEA < NiMo/RFAU < NiMo/FAU. That does not correlate
* MD – Middle distillates, ** UCO – Unconverted oil. with the expected hydrogenating activity of the catalysts according to
HRTEM and XPS data. As mentioned above, hydrogenating function of
It was found that during the first 150 h time on stream (TOS), the the catalysts expected to decrease in the following order: NiMo/
yield of the products changed, after that the yield of the products FAU ≈ NiMo/BEA ≥ NiMo/RFAU.
achieved steady state. Thus, the hydrocracking process was carried out The difference in selectivity to the middle distillates is commonly
at each temperature during 192 h TOS to obtain the data at the steady accounted for by the difference in the textural properties of the cata-
state activity of the catalysts. Table 7 summarizes the results of NiMo/ lysts, as well as the strength, concentration and availability of the BAS.
FAU, NiMo/RFAU and NiMo/BEA catalysts testing in VGO hydro- Diffusional limitations lead to secondary cracking reactions and to
cracking. formation of gas and naphtha. The decrease of crystal size of zeolites Y
In hydrocracking process, catalyst activity is the ability of a catalyst and Beta leads to the increase of selectivity to middle distillates of
to convert a specified feedstock to lighter products under specific op- hydrocracking catalysts as it was shown in works [40–42]. The average
erating conditions. In commercial hydrocracking, catalyst activity is crystal size of the BEA zeolite is smaller than that of the FAU and RFAU
expressed as a temperature required to obtain a specified conversion samples. However, selectivity to the middle distillates of the NiMo/BEA
under certain process conditions. Lower temperature is indicative of catalyst is significantly lower than the NiMo/FAU and NiMo/RFAU
higher activity [76]. Therefore, to compare the catalysts activities, it is catalysts.
suitable to use the dependence of conversion on the process tempera- Mesopore volume of the zeolites does not correlate with selectivity
ture. As can be seen in Fig. 11a, the process temperature required to of the catalysts. The highest mesopore volume is observed for RFAU
achieve 60% conversion should be 395 °C for the NiMo/BEA catalyst, zeolite, but selectivity of the NiMo/RFAU catalyst is lower than that of
400 °C for the NiMo/FAU, and 405 °C for the NiMo/RFAU. This differ- NiMo/FAU catalyst. According to pore size distributions of the zeolites
ence greatly exceeds the experimental error. Thus, the activity of the (Fig. 2) RFAU sample contains substantially higher volume of pores
catalysts increases in the following order: NiMo/RFAU < NiMo/ with diameter of 3–4 nm than FAU sample. However, zeolite FAU
FAU < NiMo/BEA. contains substantially higher volume of pores with diameter of 4–5 nm.
To explain the difference in activity, it should be noted that hy- It is unlikely that these differences can explain higher selectivity of the
drocracking catalysts are bifunctional, so their activity influenced by NiMo/FAU catalyst in comparison with the NiMo/RFAU catalyst.
acid sites and by hydrogenation sites [77–79]. The HRTEM method Probably, the difference in the selectivity of these catalysts is due to
showed that the particle size and the stacking number of sulfide par- the difference of acid sites strength and concentration. In our previous
ticles in all catalysts are similar. Also, it was shown that active com- work [39] it was shown, that the highest hydrocracking activity and
ponent particles were preferentially localized on γ-Al2O3. Therefore, middle distillates yield was achieved for NiMo/RFAU + Al2O3 catalyst
similar morphology suggests similar influence on hydrogenating ac- prepared from micro-mesoporous material characterized by optimal
tivity of the catalysts zeolitic acidity. This observation can be explained by the decrease of
In addition, XPS method showed that the Mo content in the form of the contribution of strong Brønsted acid sites (bridging Si–O(H)–Al
MoS2 is similar in all catalysts. The nickel contents in the form of NiMoS groups in faujasites cages) accompanied by the increase of the con-
phase in the NiMo/FAU and NiMo/BEA catalysts are 69 and 67%, re- tribution of slightly less acidic BAS located at the external surface of
spectively, and slightly lower in the NiMo/RFAU catalyst – 61%. zeolite microcrystallites and in the intracrystalline mesopores or in the
However, HRTEM images of the NiMo/RFAU catalyst show several bulk channels of MCM-41-like material.
sulfide particles. Thus, according to XPS and HRTEM data, it can be Increase of the proportion of strong BAS with PA = 1125–1145 kJ/
predicted that the hydrogenating function of the catalysts will be de- mol (bridging Si–O(H)–Al groups in faujasites cages, (BASin micropores)
creased in the following order: NiMo/FAU ≈ NiMo/BEA ≥ NiMo/ and decrease in the proportion of slightly less acidic BAS with
RFAU. Despite very close properties of the sulfide active component, PA = 1190–1200 kJ/mol (BASin mesopores) were observed in the fol-
the activity of NiMo/FAU and NiMo/BEA catalysts differ significantly. lowing order of the zeolites: FAU, RFAU, BEA which correlates well
The difference of hydrocracking catalysts activity is generally attributed with the respective catalysts selectivity to middle distillates. On the
to the strength and concentration of Brønsted acid sites, as well as their other hand, it can be assumed that less acidic BASin mesopores are loca-
availability. Indeed, the concentration of strong BAS, as well as the total lized primarily at the external surface of zeolite microcrystallites and/
BAS content for the zeolite component of the catalysts, in the RFAU, or in the intracrystalline mesopores or in the channels of MCM-41-like
FAU, BEA series increases, which is well correlated with the activity of material for RFAU sample and, therefore, diffusional limitations de-
the catalysts in hydrocracking of VGO. In addition, in the same order of crease and the selectivity to middle distillates increases.
the samples the average size of the zeolite crystals decreases, which Besides, the strong BASin micropores densities increase in the following
should enhance the availability of acid sites and therefore lead to in- order: FAU ≈ RFAU < BEA, which can also contribute to over-
creasing activity of the catalysts in VGO hydrocracking, as it was found cracking.
in the papers [40–42]. However, the availability of acid sites in the In the studies [17,43] dedicated to direct comparison of the

187
P.P. Dik et al. Fuel 237 (2019) 178–190

Fig. 11. Activity and selectivity of the catalysts in VGO hydrocracking: a) dependence of conversion on the temperature b) selectivity to gas as a function of
conversion; c) selectivity to naphtha as a function of conversion; d) selectivity to middle distillates as a function of conversion.

catalysts based on zeolite Y and zeolite Beta in the VGO hydrocracking, zeolite Beta (BEA) with a small particle sizes, zeolite Y obtained by
it was found that the catalysts based on zeolite Beta had a lower se- recrystallization (RFAU). According to HRTEM, low-temperature ad-
lectivity to middle distillates than the zeolite Y based catalysts. It is sorption of nitrogen, TPD-NH3 and FTIR spectroscopy of adsorbed CO,
likely that the structure of zeolite Beta micropores is less suitable to zeolites had different crystal sizes, mesopore volume, strength and
obtaine middle distillates than the structure of zeolite Y micropores. concentration of acid sites. According to HRTEM and XPS, sulfide cat-
The catalysts selectivity to gas and naphtha ranged in the reverse se- alysts have the close size and stacking number of the MoS2 particles, as
quence compared to the selectivity to middle distillates. well as NiMoS phase content and Mo sulfidation degree. The catalysts
The selectivity to gas and naphtha decreases in following order: were tested in hydrocracking of VGO. The NiMo/BEA catalyst with the
NiMo/BEA > NiMo/RFAU > NiMo/FAU (Fig. 11b,c). The explana- smallest zeolite average particle size and the highest strength and
tion of the difference of the catalysts selectivities to gas and naphtha is concentration of BAS had the highest activity and the lowest selectivity
similar to explaining the difference of the catalysts selectivity to middle to middle distillates in VGO hydrocracking. Despite the highest meso-
distillates: secondary cracking reactions increase the yield of gas and pore volume of zeolite, the NiMo/RFAU catalyst showed less activity
naphtha. It should be noted that in some cases, naphtha is desired and selectivity to middle distillates than the NiMo/FAU catalyst with
product of hydrocracking process, so highly selective to naphtha cata- the smaller average zeolite crystal size and slightly higher BAS con-
lysts are used in such cases. However the high yield of gas is anyway centration in the zeolite. The highest yield of middle distillates (57.8 wt
undesirable because of its low cost. VGO hydrocracking on the NiMo/ %) was obtained by VGO hydrocracking on the NiMo/FAU catalyst. It
BEA catalyst results in much higher gas yield than on other catalysts. was found that the selectivity to middle distillates of the studied cata-
Lewis acid sites can cause high gas yield in heavy fractions cracking lysts correlated well with the ratio of the concentration of strong BASin
[80,81] Concentration of strong LAS in the BEA sample is more than mesopores to the BASin micropores: the higher the ratio, the higher the se-
1.5–2 times higher than in FAU and RFAU zeolites Table 5. lectivity to middle distillates. Probably, it is due to both higher avail-
Possibility of the application of zeolite Beta for the preparation of ability and lower acidity of BASin mesopores in comparison with BASin
highly selective to naphtha hydrocracking catalysts requires further micropores. Using of zeolite Beta for hydrocracking catalysts preparation
study, because they suffer of high selectivity to gas. appears to result in catalysts with high selectivity to naphtha and gas.

4. Conclusions Acknowledgement

NiMo/zeolite-Al2O3 hydrocracking catalysts were prepared using This work was conducted within the framework of the budget
three different zeolites with a close Si/Al ratio: zeolite Y (FAU) and project #AAAA-A17-117041710077-4 for Boreskov Institute of

188
P.P. Dik et al. Fuel 237 (2019) 178–190

Catalysis and within the State Program of Topchiev Institute of Catal 2004;154:1411–7.
Petrochemical Synthesis RAS supported by FASO Russia. [29] Zhang Z, Liu X, Xu Y, Xu R. Realumination of dealuminated zeolites Y. Zeolites
1991;11:232–8. https://doi.org/10.1016/S0144-2449(05)80224-6.
[30] Qin Z, Shen B, Yu Z, Deng F, Zhao L, Zhou S, et al. A defect-based strategy for the
References preparation of mesoporous zeolite y for high-performance catalytic cracking. J
Catal 2013;298:102–11. https://doi.org/10.1016/j.jcat.2012.11.023.
[31] De Jong KP, Zečević J, Friedrich H, De Jongh PE, Bulut M, Van Donk S, et al. Zeolite
[1] Dahlberg AJ, Mukherjee UK. Encyclopaedia of Hydrocarbons. vol. II. 2005.
Y crystals with trimodal porosity as ideal hydrocracking catalysts. Angew Chem
[2] Speight JG. The refinery of the. Future 2011. https://doi.org/10.1016/B978-0-
2010;49:10074–8. https://doi.org/10.1002/anie.201004360.
8155-2041-2.10009-8.
[32] He J, Duan X, Evans DG, Howe RF. Mesoporous material from zeolite. J Porous
[3] Ward JW. Design and preparation of hydrocracking catalysts. Stud Surf Sci Catal
Mater 2002;9:43–8. https://doi.org/10.1023/A:1014303906232.
1983;16:587–618. https://doi.org/10.1016/S0167-2991(09)60052-5.
[33] Ivanova II, Kuznetsov AS, Yuschenko VV, Knyazeva EE. Design of composite micro/
[4] Egia B, Cambra JF, Arias PL, Güemez MB, Legarreta JA, Pawelec B, et al. Surface
mesoporous molecular sieve catalysts. Pure Appl Chem 2018;76:1647–58.
properties and hydrocracking activity of NiMo zeolite catalysts. Appl Catal A Gen
[34] Inagaki S, Ogura M, Inami T, Sasaki Y, Kikuchi E, Matsuka M. Synthesis of MCM-41-
1998;169:37–53. https://doi.org/10.1016/S0926-860X(97)00320-7.
type mesoporous materials using filtrate of alkaline dissolution of ZSM-5 zeolite.
[5] Šimáček P, Kubička D. Hydrocracking of petroleum vacuum distillate containing
Microporous Mesoporous Mater 2004;74:163–70. https://doi.org/10.1016/j.
rapeseed oil: evaluation of diesel fuel. Fuel 2010;89:1508–13. https://doi.org/10.
micromeso.2004.06.004.
1016/j.fuel.2009.09.029.
[35] Ordomsky VV, Murzin VY, Monakhova YV, Zubavichus YV, Knyazeva EE,
[6] Chop W, Lee K, Choi K, Ha B. Hydrocracking of vacuum gas oil on CoMo/alumina
Nesterenko NS, et al. Nature, strength and accessibility of acid sites in micro/me-
(or silica-alumina) containing zeolite. Stud Surf Sci Catal 1999;127:243–50.
soporous catalysts obtained by recrystallization of zeolite BEA. Microporous
[7] Zheng X, Chang J, Fu Y. One-pot catalytic hydrocracking of diesel distillate and
Mesoporous Mater 2007;105:101–10. https://doi.org/10.1016/j.micromeso.2007.
residual oil fractions obtained from bio-oil to gasoline-range hydrocarbon fuel. Fuel
05.056.
2015;157:107–14. https://doi.org/10.1016/j.fuel.2015.05.002.
[36] García-Martínez J, Johnson M, Valla J, Li K, Ying JY. Mesostructured zeolite
[8] Minja RJA, Ternan M. Hydrocracking Boscan heavy oil with a Co-Mo/Al2O3 cata-
Y—high hydrothermal stability and superior FCC catalytic performance. Catal Sci
lyst containing an H-mordenite zeolite component. Energy Fuels 1991;5:117–22.
Technol 2012;2:987–94. https://doi.org/10.1039/c2cy00309k.
https://doi.org/10.1021/ef00025a020.
[37] Ivanova II, Knyazeva EE. Micro–mesoporous materials obtained by zeoliter-
[9] Murata K, Liu Y, Watanabe MM, Inaba M, Takahara I. Hydrocracking of algae oil
ecrystallization: synthesis, characterization and catalytic applications. Chem Soc
into aviation fuel-range hydrocarbons using a Pt-Re catalyst. Energy Fuels
Rev 2013;42:3671–88. https://doi.org/10.1039/C2CS35341E.
2014;28:6999–7006. https://doi.org/10.1021/ef5018994.
[38] Ivanova II, Kasyanov IA, Maerle AA, Zaikovskii VI. Mechanistic study of zeolites
[10] Gutiérrez A, Arandes JM, Castaño P, Olazar M, Bilbao J. Preliminary studies on fuel
recrystallization into micro-mesoporous materials. Microporous Mesoporous Mater
production through LCO hydrocracking on noble-metal supported catalysts. Fuel
2014;189:163–72. https://doi.org/10.1016/j.micromeso.2013.11.001.
2012;94:504–15. https://doi.org/10.1016/j.fuel.2011.10.010.
[39] Kazakov MO, Nadeina KA, Danilova IG, Dik PP, Klimov OV, Pereyma VY, et al.
[11] Dik PP, Klimov OV, Danilova IG, Leonova KA, Pereyma VY, Budukva SV, et al.
Hydrocracking of vacuum gas oil over NiMo/Y-Al2O3: effect of mesoporosity in-
Hydroprocessing of hydrocracker bottom on Pd containing bifunctional catalysts.
troduced by zeolite Y recrystallization. Catal Today 2018;305:117–25. https://doi.
Catal Today 2016;271:154–62. https://doi.org/10.1016/j.cattod.2015.09.050.
org/10.1016/j.cattod.2017.08.048.
[12] Lee J, Hwang S, Lee SB, Song IK. Production of middle distillate through hydro-
[40] Camblor MA, Corma A, Martínez A, Martínez-Soria V, Valencia S. Mild hydro-
cracking of paraffin wax over NiMo/TiO2-SiO2 catalysts. Korean J Chem Eng
cracking of vacuum gasoil over NiMo-beta zeolite catalysts: the role of the location
2010;27:1755–9. https://doi.org/10.1007/s11814-010-0279-3.
of the NiMo phases and the crystallite size of the zeolite. J Catal 1998;179:537–47.
[13] Hita I, Cordero-Lanzac T, Gallardo A, Arandes JM, Rodríguez-Mirasol J, Bilbao J,
https://doi.org/10.1006/jcat.1998.2249.
et al. Phosphorus-containing activated carbon as acid support in a bifunctional Pt-
[41] Landau MV, Vradman L, Valtchev V, Lezervant J, Liubich E, Talianker M.
Pd catalyst for tire oil hydrocracking. Catal Commun 2016;78:48–51. https://doi.
Hydrocracking of heavy vacuum gas oil with a Pt/H-beta–Al2O3 catalyst: effect of
org/10.1016/j.catcom.2016.01.035.
zeolite crystal size in the nanoscale range. Ind Eng Chem Res 2003;42:2773–82.
[14] Gutiérrez A, Arandes JM, Castaño P, Olazar M, Bilbao J. Effect of pressure on the
https://doi.org/10.1021/ie020899o.
hydrocracking of light cycle oil with a Pt-Pd/HY catalyst. Energy Fuels
[42] Cui Q, Zhou Y, Wei Q, Tao X, Yu G, Wang Y, et al. Role of the zeolite crystallite size
2012;26:5897–904. https://doi.org/10.1021/ef3009597.
on hydrocracking of vacuum gas oil over NiW/Y-ASA catalysts. Energy Fuels
[15] Dik PP, Klimov OV, Koryakina GI, Leonova KA, Pereyma VY, Budukva SV, et al.
2012;26:4664–70.
Composition of stacked bed for VGO hydrocracking with maximum diesel yield.
[43] Hassan A, Ahmed S, Ali MA, Hamid H, Inui T. A comparison between β- and USY-
Catal Today 2014;220–222:124–32. https://doi.org/10.1016/j.cattod.2013.07.
zeolite-based hydrocracking catalysts. Appl Catal A Gen 2001;220:59–68. https://
004.
doi.org/10.1016/S0926-860X(01)00705-0.
[16] Yan TY. Modified zeolite-based catalyst for effective extinction hydrocracking. Ind
[44] Ivanova AS, Korneeva EV, Bukhtiyarova GA, Nuzhdin AL, Budneva AA, Prosvirin IP,
Eng Chem Res 1989;28:1463–6. https://doi.org/10.1021/ie00094a004.
et al. Hydrocracking of vacuum gas oil in the presence of supported nickel-tungsten
[17] Ali MA, Tatsumi T, Masuda T. Development of heavy oil hydrocracking catalysts
catalysts. Kinet Catal 2011;52:446–58. https://doi.org/10.1134/
using amorphous silica-alumina and zeolites as catalyst supports. Appl Catal A Gen
S0023158411030098.
2002;233:77–90. https://doi.org/10.1016/S0926-860X(02)00121-7.
[45] Ishihara A, Itoh T, Nasu H, Hashimoto T, Doi T. Hydrocracking of 1-methyl-
[18] Corma A, Martinez A, Pergher S, Peratello S, Perego C, Bellusi G. Hydrocracking-
naphthalene/decahydronaphthalene mixture catalyzed by zeolite-alumina compo-
hydroisomerization of n-decane on amorphous silica-alumina with uniform pore
site supported NiMo catalysts. Fuel Process Technol 2013;116:222–7. https://doi.
diameter. Appl Catal A Gen 1997;152:107–25. https://doi.org/10.1016/S0926-
org/10.1016/j.fuproc.2013.07.001.
860X(96)00338-9.
[46] Chen S, Yang Y, Zhang K, Wang J. BETA zeolite made from mesoporous material
[19] Leyva C, Rana MS, Trejo F, Ancheyta J. On the use of acid-base-supported catalysts
and its hydrocracking performance. Catal Today 2006;116:2–5. https://doi.org/10.
for hydroprocessing of heavy petroleum. Ind Eng Chem Res 2007;46:7448–66.
1016/j.cattod.2006.04.005.
https://doi.org/10.1021/ie070128q.
[47] Jiang J, Dong Z, Chen H, Sun J, Yang C, Cao F. The effect of additional zeolites in
[20] Wade R, Vislosky J, Maesen T, Torchia D. Improvements to hydrocracking catalyst
amorphous silica – alumina supports on hydrocracking of semirefined paraffinic
activity and selectivity at various operational and feedstocks conditions are dis-
wax 2013:1035–9.
cussed. Pet Technol Q 2009;3:1.
[48] Ivanova II, Knyazeva EE, Dobryakova IV, Monakhova YM, Kozhina OV, Tikhonova
[21] Marafi M, Stanislaus A, Furimsky E. Handbook of spent hydroprocessing catalysts.
AA. Patent RU. No. 2393992, 2009.
2017. doi: 10.1016/B978-0-444-63881-6.00004-4.
[49] Pashigreva AV, Klimov OV, Bukhtiyarova GA, Fedotov MA, Kochubey DI, Chesalov
[22] Manrique C, Guzmán A, Pérez-Pariente J, Márquez-Álvarez C, Echavarría A.
YA, et al. The superior activity of the CoMo hydrotreating catalysts, prepared using
Vacuum gas-oil hydrocracking performance of Beta zeolite obtained by hydro-
citric acid: What’s the reason? vol. 175. Elsevier Masson SAS; 2010. doi: 10.1016/
thermal synthesis using carbon nanotubes as mesoporous template. Fuel
S0167-2991(10)75014-X.
2016;182:236–47. https://doi.org/10.1016/j.fuel.2016.05.097.
[50] Jos Van Dillen A, Terörde RJAM, Lensveld DJ, Geus JW, De Jong KP. Synthesis of
[23] Ding L, Zheng Y, Zhang Z, Ring Z, Chen J. HDS, HDN, HDA, and hydrocracking of
supported catalysts by impregnation and drying using aqueous chelated metal
model compounds over Mo-Ni catalysts with various acidities. Appl Catal A Gen
complexes. J Catal 2003;216:257–64. https://doi.org/10.1016/S0021-9517(02)
2007;319:25–37. https://doi.org/10.1016/j.apcata.2006.11.016.
00130-6.
[24] Vogt ETC, Gareth WT, Chowdhury AD, Weckhuysen BM. Zeolites and zeotypes for
[51] Klimov OV, Nadeina KA, Dik PP, Koryakina GI, Pereyma VY, Kazakov MO, et al.
oil and gas conversion. Adv Catal 2015;58:143–314. https://doi.org/10.1016/bs.
CoNiMo/Al2O3catalysts for deep hydrotreatment of vacuum gasoil. Catal Today
acat.2015.10.001.
2016;271:56–63. https://doi.org/10.1016/j.cattod.2015.11.004.
[25] Corma A. State of the art and future challenges of zeolites as catalysts. J Catal
[52] Pereyma VY, Dik PP, Klimov OV, Budukva SV, Leonova KA, Noskov AS.
2003;216:298–312. https://doi.org/10.1016/S0021-9517(02)00132-X.
Hydrocracking of vacuum gas oil in the presence of catalysts NiMo/
[26] Szostak R. Chapter 6 Secondary synthesis methods. Stud Surf Sci Catal
Al2O3–amorphous aluminosilicates and NiW/Al2O3–amorphous aluminosilicates.
2001;137:261–97. https://doi.org/10.1016/S0167-2991(01)80248-2.
Russ J Appl Chem 2015;88:1969–75. https://doi.org/10.1134/
[27] Agudelo JL, Mezari B, Hensen EJM, Giraldo SA, Hoyos LJ. On the effect of EDTA
S10704272150120113.
treatment on the acidic properties of USY zeolite and its performance in vacuum gas
[53] Hadjiivanov K. Identification and characterization of surface hydroxyl groups by
oil hydrocracking. Appl Catal A Gen 2014;488:219–30. https://doi.org/10.1016/j.
infrared spectroscopy. Adv Catal 2014;57:99–318. https://doi.org/10.1016/B978-
apcata.2014.10.007.
0-12-800127-1.00002-3.
[28] Lutz W, Rüscher C, Gesing T. Investigations of the mechanism of dealumination of
[54] Toktarev AV, Malysheva LV, Paukshtis EA. Effect of thermal treatment tonditions
zeolite Y by steam: tuned mesopore formation versus the Si/Al ratio. Stud Surf Sci
on the acid properties of zeolite beta. Kinet Catal 2010;51:318–24. https://doi.org/

189
P.P. Dik et al. Fuel 237 (2019) 178–190

10.1134/S0023158410020229. [69] Galtayries A, Wisniewski S, Grimblot J. Formation of thin oxide and sulphide films
[55] Paukshtis EA, Yurchenko EN. Study of the acid-base properties of heterogeneous on polycrystalline molybdenum foils: characterization by XPS and surface potential
catalysts by infrared spectroscopy. Russ Chem Rev 1983;52:242–58. https://doi. variations. J Electron Spectros Relat Phenomena 1997;87:31–44. https://doi.org/
org/10.1070/RC1983v052n03ABEH002812. 10.1016/S0368-2048(97)00071-6.
[56] Cairon O, Chevreau T, Lavalley J. Bronsted acidity of extraframework debris in [70] Barrera MC, Viniegra M, Escobar J, Vrinat M, De Los Reyes JA, Murrieta F, et al.
steamed Y zeolites from the FTIR study of CO adsorption. J Chem Soc Highly active MoS2 on wide-pore ZrO2-TiO2 mixed oxides. Catal Today
1998;94:3039–47. https://doi.org/10.1039/A803886D. 2004;98:131–9. https://doi.org/10.1016/j.cattod.2004.07.027.
[57] Soltanov RI, Paukshtis EA, Yurchenko EN. IR spectroscopic investigation of the [71] Okamoto Y, Toshinobu I, Shiichiro T. Surface structure of CoO-MoO3/Al2O3 cat-
thermodynamics of the reaction of carbon-monoxide with the surface of several alysts studied by X-ray photoelectron spectroscopy. J Catal 1980;65:448–60.
oxide adsorbents. Kinet Catal 1982;23:135–41. https://doi.org/10.1016/0021-9517(80)90322-X.
[58] Bej SK, Dalai AK, Maity SK. Effect of diluent size on the performance of a micro- [72] Alstrup I, Chorkendorff I, Candia R, Clausen BS, Topsøe H. A combined X-ray
scale fixed bed multiphase reactor in up flow and down flow modes of operation. photoelectron and Mössbauer emission spectroscopy study of the state of cobalt in
Catal Today 2001;64:333–45. https://doi.org/10.1016/S0920-5861(00)00536-8. sulfided, supported, and unsupported CoMo catalysts. J Catal 1982;77:397–409.
[59] Francis J, Guillon E, Bats N, Pichon C, Corma A, Simon LJ. Design of improved https://doi.org/10.1016/0021-9517(82)90181-6.
hydrocracking catalysts by increasing the proximity between acid and metallic sites. [73] Ninh TKT, Massin L, Laurenti D, Vrinat M. A new approach in the evaluation of the
Appl Catal A Gen 2011;409–410:140–7. https://doi.org/10.1016/j.apcata.2011.09. support effect for NiMo hydrodesulfurization catalysts. Appl Catal A Gen
040. 2011;407:29–39. https://doi.org/10.1016/j.apcata.2011.08.019.
[60] Manrique C, Guzmán A, Pérez-Pariente J, Márquez-Álvarez C, Echavarría A. Effect [74] Mozhaev AV, Nikulshin PA, Pimerzin AA, Maslakov KI, Pimerzin AA. Investigation
of synthesis conditions on zeolite Beta properties and its performance in vacuum gas of co-promotion effect in NiCoMoS/Al2O3catalysts based on Co2Mo10-hetero-
oil hydrocracking activity. Microporous Mesoporous Mater 2016;234:347–60. polyacid and nickel citrate. Catal Today 2016;271:80–90. https://doi.org/10.1016/
https://doi.org/10.1016/j.micromeso.2016.07.017. j.cattod.2015.11.002.
[61] Zecchina A, Bordiga S, Spoto G, Scarano D, Petrini G, Leofanti G, et al. Interaction of [75] Coulier L, De Beer VHJ, Van Veen JAR, Niemantsverdriet JW. Correlation between
CO with nanosized ZSM5 and silicalite. J Chem Soc Faraday Trans hydrodesulfurization activity and order of Ni and Mo sulfidation in planar silica-
1992;88:2959–69. supported NiMo catalysts: the influence of chelating agents. J Catal
[62] Kiricsi I, Flego C, Pazzuconi G, Parker Jr. WO, Millini R, Perego C, et al. Progress 2001;197:26–33. https://doi.org/10.1006/jcat.2000.3068.
toward understanding zeolite β acidity: an IR and 27Al NMR spectroscopic study. J [76] Scherzer J, Gruia AJ. Hydrocracking science and technology. CRC Press; 1996.
Phys Chem 1994;98:4627–34. https://doi.org/10.1021/j100068a024. [77] Alsobaai AM, Zakaria R, Hameed BH. Gas oil hydrocracking on NiW/USY catalyst:
[63] Corma A, Forn V, Melo F. Zeolite beta : structure, activity, and selectivity for cat- effect of tungsten and nickel loading. Chem Eng J 2007;132:77–83. https://doi.org/
alytic cracking. ACS Symp Ser 1988;375:49–63. 10.1016/j.cej.2007.01.021.
[64] Daniell W, Topsøe NY, Knözinger H. An FTIR study of the surface acidity of USY [78] Henry R, Tayakout-Fayolle M, Afanasiev P, Lorentz C, Lapisardi G, Pirngruber G.
zeolites: comparison of CO, CD3CN, and C5H5N probe molecules. Langmuir Vacuum gas oil hydrocracking performance of bifunctional Mo/Y zeolite catalysts
2001;17:6233–9. https://doi.org/10.1021/la010345a. in a semi-batch reactor. Catal Today 2014;220–222:159–67. https://doi.org/10.
[65] Montanari T, Finocchio E, Busca G. Infrared spectroscopy of heterogeneous cata- 1016/j.cattod.2013.06.024.
lysts: acidity and accessibility of acid sites of faujasite-type solid acids. J Phys Chem [79] Gutberlet LC, Bertolacini RJ, Kukes SG. Design of a nickel-tungsten hydrocracking
C 2011;115:937–43. https://doi.org/10.1021/jp103567g. catalyst. Energy Fuels 1994;8:227–33. https://doi.org/10.1021/ef00043a035.
[66] Anderson MW, Klinowski J. Zeolites treated with silicon tetrachloride vapour. IV. [80] Corma A, García H. Lewis acids: from conventional homogeneous to green homo-
Acidity. Zeolites 1986;6:455–66. https://doi.org/10.1016/0144-2449(86)90030-8. geneous and heterogeneous catalysis. Chem Rev 2003;103:4307–65. https://doi.
[67] Gabrienko AA, Danilova IG, Arzumanov SS, Toktarev AV, Freude D, Stepanov AG. org/10.1021/cr030680z.
Strong acidity of silanol groups of zeolite beta: evidence from the studies by IR [81] Schallmoser S, Ikuno T, Wagenhofer MF, Kolvenbach R, Haller GL, Sanchez-Sanchez
spectroscopy of adsorbed CO and 1H MAS NMR. Microporous Mesoporous Mater M, et al. Impact of the local environment of Brønsted acid sites in ZSM-5 on the
2010;131:210–6. https://doi.org/10.1016/j.micromeso.2009.12.025. catalytic activity in n-pentane cracking. J Catal 2014;316:93–102. https://doi.org/
[68] Jentys A, Pham NH, Vinek H. Nature of hydroxy groups in MCM-41. J Chem Soc 10.1016/j.jcat.2014.05.004.
Faraday Trans 1996;92:3287. https://doi.org/10.1039/ft9969203287.

190

You might also like