Preparation of Mesoporous Zeolite Y by Fluorine–Alkaline Treatment

Apr 19, 2019 - Lanzhou Petrochemical Center, Petrochemical Research Institute, ... Facile Fabrication of Mesoporous Hierarchical Co-Doped ZnO for High...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IECR

Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Preparation of Mesoporous Zeolite Y by Fluorine−Alkaline Treatment for Hydrocracking Reaction of Naphthalene Shenyong Ren,*,† Bo Meng,† Xiao Sui,† Hongchang Duan,‡ Xionghou Gao,‡ Haitao Zhang,‡ Penghui Zeng,† Qiaoxia Guo,† and Baojian Shen*,† †

Ind. Eng. Chem. Res. Downloaded from pubs.acs.org by UNIV AUTONOMA DE COAHUILA on 04/26/19. For personal use only.

State Key Laboratory of Heavy Oil Processing, the Key Laboratory of Catalysis of CNPC, College of Chemical Engineering and Environment, College of Sciences, China University of Petroleum, 18# Fuxue Road, Changping, Beijing 102249, China ‡ Lanzhou Petrochemical Center, Petrochemical Research Institute, PetroChina Company Limited, Lanzhou 730060, China ABSTRACT: In this study mesoporous zeolite Y (USY-F-AT) was prepared by fluorine−alkaline combined treatment method. The properties of USY-F-AT were characterized by XRD, N2 physisorption, NH3-TPD, XPS, and TEM. USY-F-AT provided greater mesopore volume and surface area than conventional USY. NH3TPD showed the USY-F-AT has many more acid sites and stronger acid strength compared with USY. Ni and Mo were impregnated into USY and USY-F-AT to form catalysts NiMo/(Al2O3 + USY) and NiMo/(Al2O3 + USY-F-AT), respectively. The hydrogenation activity and cracking ability of catalysts NiMo/ (Al2O3 + USY) and NiMo/(Al2O3 + USY-F-AT) were studied by hydrocracking reaction using naphthalene as model compound. The results showed that catalyst NiMo/(Al2O3 + USY-F-AT) exhibited similar hydrogenation activity but much higher ring-opening ability in comparison with catalyst NiMo/(Al2O3 + USY).

compared with other materials in cracking catalysts.11−15 However, zeolite Y is a microporous material with disadvantages that active site is poorly accessible and diffusion is limited.3,4 Therefore, many modifications of zeolite Y have been developed to enhance the catalytic behavior.16−18 It was reported that desilication is an effective method for the preparation of mesoporous zeolites.19−23 An optimal treatment range of Si/Al molar ratio is 25−50 for desilication.24 The fluorine−alkaline combined treatmentmodified ZSM-5 (Si/Al = 14.3) can dramatically increase the mesoporous surface area and mesoporous volume.25 In this work, the fluorine−alkaline combined treatment-modified Y can ensure high mesopore volume and appropriate acid sites for hydrocracking performance of catalysts. The hydrocracking reactions of mesoporous Y and corresponding catalysts NiMo/ (Al2O3 + USY) and NiMo/(Al2O3 + USY-F-AT) were studied through a microreactor using naphthalene as a feedstock to expose their hydrogenation activity and ring-opening performance.

1. INTRODUCTION Increasingly stringent environmental regulations indicate future clean diesel needs to be in higher demand: lower sulfur and aromatic content and higher cetane number. However, light cycle oil (LCO) is rich in polycyclic aromatic hydrocarbons, which resulted in low diesel cetane number and negative environmental impact.1 Only hydrogenation (e.g., naphthalene hydrogenation to form tetralin and decalin) can slightly increase its cetane number, and monocyclic cycloalkanes with side chains and the normal paraffins formed from ring-opening reaction of polycyclic aromatics can significantly increase the cetane number. The FCC diesel oil upgrading catalyst must have high selective ring opening and control aromatic hydrogenation saturation ability, while avoiding cracking of alkyl side chains and isomerization and cleavage of alkanes. However, catalyst technology to guarantee the high activity of diesel hydrogenation and selectivity of ring opening is very difficult. Traditionally, the majority of experimental projects have focused on deep hydrotreating and hydrocracking to improve cetane number,2,3 but both of them have key drawbacks; high diesel fuel yield and high cetane number cannot be afforded at the same time. By contrast, deep hydrogenation and selective ring opening can yield high octane number of diesel fuel. As a result, FCC diesel fuel refining catalyst must provide high aromatic hydrogenation saturation, hydrocracking, and selective ring-opening (SRO) ability.4−6 Several works show that adding appropriate acidic constituents into the conventional Al2O3-supported hydrocracking catalysts can substantially improve the quality of the diesel fuel. Zeolite Y is considered as the most efficient acidic component due to its high activity and selectivity in middle distillates7−10 © XXXX American Chemical Society

2. EXPERIMENTAL SECTION 2.1. Preparation of Modified Y Zeolite and Composite Supports of Catalytic Reaction. The USY was prepared using NaY as starting material (Si/Al = 2.85, Lanzhou Petrochemical Company) by ammonium exchange, Received: Revised: Accepted: Published: A

January 23, 2019 April 9, 2019 April 19, 2019 April 19, 2019 DOI: 10.1021/acs.iecr.9b00422 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

were measured after desorption of pyridine at 200 °C for 1 h and at 350 °C for 1 h under vacuum. Transmission electron microscope (TEM) pictures were taken using JEO 2100FX transmission microscope operated at 200 kv. X-ray photoelectron spectroscopy (XPS) study was performed with a Thermo Fisher K-Alpha system using Al Kα (hv = 1486.6 eV). 2.3. Catalytic Performance Evaluation. The hydrocracking reaction of naphthalene was carried out in a stainless steel fixed bed reactor (JQ-II hydro-processing unit). The system included a high-pressure reactor (internal diameter 7 mm, length 28 cm). Two thermocouples which were placed at two fixed sites of the upper and lower furnace, respectively, controlled the reaction temperature. The temperature of reactor was measured by a thermocouple which was placed in the catalyst bed center. H2 gas entered into the reactor through a manometer and a gas flowmeter; reactant was pumped into the reactor by two lines. Reaction conditions were as follows: model feed 10 wt % naphthalene and catalyst, loading 1 g, temperatures 380 and 400 °C, respectively, gas H2 pressure 4 MPa, w8 hly space velocity (WHSV) 8 h−1, and H2/ oil ratio 500 (v/v). The presulfidation conditions for catalysts wre as follows: presulfidation solution 10 wt % CS2 in decane, temperature 330 °C, at 4 MPa for 4 h.

followed by 100% steam treating at 650 °C for 2 h. The Na content of USY is 0.5%. The process of fluorination−alkaline treatment: the aforementioned USY (30 g) was first stirred in distilled water, followed by addition of ammonium fluoride (NH4F, Sinopharm Chemical Reagent Co., Ltd., China) solution (0.75 mol/L) drop by drop to the configured sample. The mixed material was stirred (300 rpm) at 90 °C for 8 h. After filtration the sample was dried at 120 °C overnight and then submitted to a steaming treatment at 550 °C for 2 h. Then the obtained sample was treated by sodium hydrate (NaOH, Tianjin Guangfu Fing Chemical Research Institute) solution (0.25 mol/L). The alkaline-treated sample was collected by washing and dried at 120 °C overnight. Finally, the treated zeolite was prepared by ammonium exchange and subsequent calcination at 550 °C for 2 h, which was named USY-F-AT. The Na content of USY-F-AT is 1.3%. The support was prepared as follows: (1) 6.0 g of USY or modified USY-F-AT, 30.0 g of pseudoboehmit, and 0.72 g of sesbania power were mechanically mixed; (2) 0.72 g of nitric acid, 0.72 g of citric acid, and 16.81 g of deionized water were added to the sample. Then the mixture was kneaded and extruded to form a clover-shaped particle, which was dried at room temperature and 120 °C and then calcined at 550 °C for 4 h. Ni and Mo were introduced into this zeolite by the incipient wetness impregnation method using nickel nitrate (Ni(NO3)2·6H2O, Sinopharm Chemical Reagent Co., Ltd., Ch i n a ) a n d a m m on i u m m ol y b d at e t e t r ah y d r a t e ((NH4)6Mo7O24·4H2O, Tianjin Guangfu Fing Chemical Research Institute) solution at a metal loading of 5 and 15 wt %, respectively. The impregnated samples were calcined at 550 °C for 4 h to form catalysts NiMo/(Al2O3 + USY) and NiMo/(Al2O3 + USY-F-AT). 2.2. Characterization. X-ray diffraction (XRD) patterns of the samples were obtained on a PANalytial X’Pert Power X-ray diffractometer with monochromatized Cu Kα radiation (40Kv, 40 mA). Nitrogen physisorption measurement was performed on a Micromeritics ASAP 3020 apparatus. Before the measurement, samples were degassed at 350 °C for 8 h. With BET (Brunauer, Emmett, and Teller) method, the total surface area (SBET) of samples was measured. The external surface area, micropore surface area, and micropore volume were calculated by the tplot method. The total pore volumes (Vp) were measured at volume of N2 adsorbed at P/P0 = 0.99. The acid amounts of the samples were measured by temperature-programmed desorption of ammonia (NH3TPD) using a Micromeritics Autochem 2920 instrument. During the analysis, the samples were degassed under the condition of He gas purging with heating (from room temperature to 600 °C, 15 °C min−1) and maintained for 30 min at 600 °C. The NH3 was adsorbed by the samples after cooling down to below 100 °C. The temperature was raised to 100 °C, and then the chemical desorption curves of ammonia were recorded from 100 to 600 °C (10 °C min−1). Acid type distributions and acid sites of the samples were measured by infrared spectroscopy of chemisorbed pyridine (Py-IR) on a thermo Fisher Nicolet IS10 FT-IR spectrometer. The samples were put in a sample pool connected to the vacuum system. Then the samples were degassed at 400 °C for 2 h in a vacuum (10−6−10−7 mbar). When the temperature was cooled down to room temperature, the samples absorbed the pyridine to saturation (15 min). The FT-IR of the samples

3. RESULTS AND DISCUSSION 3.1. Characterization of USY Zeolites. 3.1.1. XRD. Figure 1 shows XRD patterns of the USY and USY-F-AT

Figure 1. XRD patterns of USY and USY-F-AT.

samples between 5° and 35°. USY-F-AT showed typical peaks of Y zeolite which indicated the good retaining of framework structure of zeolite after fluorine−alkaline treating. The relative crystallinities of the samples were 85% (USY) and 57% (USYF-AT), respectively; after fluorine−alkaline treatment, the stucture of zeolite Y has been partially destroyed. 3.1.2. Textural Properties. The textures of samples are shown in Table 1. The fluorine−alkaline treated samples possessed textural characteristics different from their parent. As shown in Figure 2, the isotherm of the USY-F-AT presented a pronounced hysteresis loop at the relative pressure of P/P0 = 0.4−0.9 which indicated mesopores formation in the sample. It could be confirmed from the pore size distribution curve that the mesopore diameter is around 7 nm (Figure 3). The B

DOI: 10.1021/acs.iecr.9b00422 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

concentration during fluorine treatment could significantly enhance the mesopore creating. 3.1.3. TEM. To further verify the mesopore formation on fluorine−alkaline-treated zeolite, TEM was employed to characterize the material. Mesopores can be obviously observed on samples USY-F-AT as shown in Figure 4. The

Table 1. Framework Composition and the Textual Properties of USY and USY-F-AT samples USY USY-FAT

SBETa (m2 g−1)

Smicrob (m2 g−1)

Smesoc (m2 g−1)

Vpored (cm3 g−1)

Vmicrob (cm3 g−1)

Vmesoe (cm3 g−1)

Si/Al ratiof

589 585

512 376

77 209

0.40 0.60

0.25 0.20

0.15 0.40

7.5 5.8

a Brunauer−Emmett−Teller (BET) method. bt-plot method. cSBET − Smicro. dVolume of N2 adsorbed at P/P0 = 0.99. eVpore − Vmicro. f Measured by X-ray diffraction (XRD).

Figure 4. TEM image of the zeolites.

TEM images of USY and USY-F-AT indicated that fluorine− alkalin- treated samples are rich in mesoporous and channel connectivity better than normal USY (Figure. 4). Furthermore, higher concentration fluorine treatment created more channels which are conducive to the diffusion of macromolecules and contact to more active sites. 3.1.4. Acidity Properties. The acidity of samples was studied by NH3-TPD and Py-IR. The NH3-TPD profiles of the modified zeolites are depicted in Figure 5. By contract, USY-F-

Figure 2. N2 adsorption−desorption isotherms for USY and USY-FAT.

Figure 5. NH3-TPD of the USY and USY-F-AT.

AT show higher total acid content than USY. The fact is that Al−F species is easily removed in the steaming treatment to form defective sites, i.e., silanol nests in the crystals,26 which is beneficial to mesopore formation by selective dissolution of framework Si preferentially. The increase of total acid amount is related to the lower Si/Al for USY-F-AT as shown in Table 1. The lower Si/Al ratio is closely associated with the improved extraction of Si due to the dislodged Al by fluorination. Table 2 shows the acidity distribution of samples such that USY-F-AT provided more acid content, especially B acid and strong acid. This is similar to the result of Si/Al ratio (Table 1). 3.2. Characterization of the Catalysts. 3.2.1. XRD and Textural Properties. For the catalysts, the characteristic diffraction signals of zeolite Y and Al2O3 (Figure 6) can be observed without obvious diffraction peaks of MoO3 and NiO, indicating that molybdenum and nickel oxide species dispersed well on the support.

Figure 3. Pore size distribution for the USY and USY-F-AT.

mesoporous distribution became “bimodal” with the increase of the concentration. The uniform contributions centered of USY-F-AT were around 7 and 42 nm. The pore size distribution of USY is only visible around 22 nm. Such comparison indicates the fluorination−alkaline treatment plays an important role in the “creaction of bimodal mesopore system”. Table 1 summarizes results of N2 adsorption and desorption analysis. Although their micropore volumes were slightly decreased, their external surface areas were increased obviously (Smeso; from 77 to 209 m2 g−1), a result of a distinct increasing of the volume of mesopores observed (Vmeso; from 0.15 to 0.40 cm 3 g −1 ). On the other hand, high F − C

DOI: 10.1021/acs.iecr.9b00422 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

USY). It is similar to the zeolite materials. Furthermore, fluorination−alkaline treatment could increase the strength of the strong acid as shown in Figure 5 and Figure 7. 3.2.3. XPS and TEM. Figure 8 displays the Ni 2p spectra of the sulfide catalyst. The binding energy is 856 eV attributed to

Table 2. Concentration of Brønsted (B) and Lewis (L) Acid Sites of Zeolites from IR Spectroscopy of Adsorbed Pyridine 200 °C, μmol g−1

350 °C, μmol g−1

samples

B

L

B+L

B

L

B+L

USY USY-F-AT

72.3 186.6

261.2 238.3

333.4 424.9

45.8 103.3

102.1 187.3

147.9 290.6

Figure 8. Ni 2p XPS spectra of the sulfide catalysts.

Figure 6. XRD patterns of catalysts.

The textures of the catalysts are depicted in Table 3. Among the three catalysts, no big difference of the surface area (SBET) Table 3. Textural Properties of Catalysts samples NiMo/(Al2O3 + USY) NiMo/(Al2O3 + USY-F-AT)

SBET (m2 g−1)

Smeso (m2 g−1)

Vpore (cm3 g−1)

Vmicro (cm3 g−1)

Vmeso (cm3 g−1)

238

164

0.24

0.04

0.20

213

162

0.25

0.02

0.23

and pore volume (Vpore) was observed due to the coverage of the basis pseudoboehmit carrier on the catalyst. 3.2.2. Acidity. The NH3-TPD profiles of the catalysts are shown in Figure 7. The order of total acid sites of the catalysts as is follows: NiMo/(Al2O3 + USY-F-AT) > NiMo/(Al2O3 +

Figure 9. Mo 3d XPS spectra of the sulfide catalysts.

Ni 2p3/2. Figure 9 shows the sulfide catalysts Mo 3d spectra, with a binding energy of 232.2 and 228.6 eV, belonging to Mo 3d3/2 and Mo 3d5/2, respectively. The surface atomic ratios of the sulfided catalysts were determined by XPS and are given in Table 4. Although the same metal contents are loaded over the two catalysts, the surface Mo and Ni ratios are different due to the different surface properties of the supports. It can be found that the content of surface Mo species is slightly more on NiMo/(Al2O3 + USY-F-AT) than that of NiMo/(Al2O3 + USY). The TEM images of the sulfide catalysts are given in Figure 10. MoS2 slabs are evidently observed as multilayers in the catalysts. Higher stacking numbers of NiMoS phase slabs suggested higher hydrogenation activity for aromatic hydrocarbon.27 As shown in Table 4, although the average length of MoS2 slab of catalyst NiMo/(Al2O3 + USY-F-AT) is shorter than NiMo/(Al2O3 + USY), the average layer number of MoS2

Figure 7. NH3-TPD of the catalysts. D

DOI: 10.1021/acs.iecr.9b00422 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

High number of acid sites can lead to overcracking to form small molecular alkanes. The reduced acidity suppresses the sequential ring-opening and cracking and results in the higher yield to hydrogenated tetralin. In contrast, lower hydrogenation activity with the enhanced acidity sites may result from large cracking of tetralin and decalin. The content of acid sites is the main factor that influences the hydrocracking of naphthalene. For NiMo/(Al2O3 + USY-F-AT), higher hydrogenation activity and appropriate acid sites allowed deep hydrogenation and good ring-opening yield but limited overcracking.

Table 4. Binding Energies and Surface Atomic Ratios on the Sulfide Catalyst items binding energy (eV)

surface species atomic ratio

avg no. of layers avg slab length (nm)

Mo 3d3/2 Mo 3d5/2 Ni 2p1/2 Ni 2p3/2 Mo/(Si + Al) Ni/(Si + Al) Mo/Ni 1.8 4.6

NiMo/ (Al2O3 + USY)

NiMo/ (Al2O3 + USY-F-AT)

232.20 228.60 873.90 855.60 0.068 0.043 1.581 2.4 4.1

231.58 228.28 874.40 856.50 0.071 0.044 1.614

4. CONCLUSION The fluorination−alkaline treatment-modified USY zeolites are used as the acidic constituent of the hydrocracking catalyst supports. Compared to USY, the fluorination−alkaline modification can increase the volume of mesopore and pore diameter. Furthermore, the fluorination−alkaline treatment can increase acid sites, especially strong acid and B acid. Consequently, catalyst NiMo/(Al2O3 + USY-F-AT) shows higher yield of selective ring-opening product by 18.5 wt % in hydrocracking reaction of naphthalene at 400 °C compared to NiMo/(Al2O3 + USY), which is attributed to the suitable acid content and acid distribution. The catalyst could reduce polycyclic aromatics content in the potential industrial application.

Figure 10. TEM images of the sulfide catalysts.

slab is much higher which may afford higher hydrogenation activity. 3.3. Catalytic Activity Evaluation. The hydrocracking reaction results of naphthalene are summarized in Table 5.



*Tel./Fax: +86 10 89732316. E-mail: [email protected].

Table 5. Naphthalene Conversion and Product Yields of the Catalysts samples temp (°C) conversion (%) product yields (wt %) tetralin tecalins trans-decalin cis-decalin ROP naphthalene

NiMo/(Al2O3 + USY)

ORCID

Shenyong Ren: 0000-0002-7708-9264 Baojian Shen: 0000-0003-1889-309X

NiMo/(Al2O3 + USY-FAT)

380 91.0

400 83.1

380 89.2

400 83.7

73.7 9.9 7.1 2.8 7.4 9.0

68.8 6.7 5.4 1.3 7.6 16.9

62.9 7.8 7.1 0.7 18.5 10.8

51.3 6.1 5.9 0.2 26.2 16.3

AUTHOR INFORMATION

Corresponding Author

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by Science Foundation of China University of Petroleum, Beijing (No.2462015YQ0313) and PetroChina.



REFERENCES

(1) Corma, A.; González-Alfaro, V.; Orchillés, A. V. Decalin and Tetralin as Probe Molecules for Cracking and Hydrotreating the Light Cycle Oil. J. Catal. 2001, 200 (1), 34−44. (2) Knottenbelt, C. Mossgas “gas-to-liquid” diesel fuels-an environmentally friendly option. Catal. Today 2002, 71 (3), 437−445. (3) Song, C.; Ma, X. New design approaches to ultra-clean diesel fuels by deep desulfurization and deep dearomatization. Appl. Catal., B 2003, 41 (1), 207−238. (4) Cooper, B. H.; Donnis, B. B. L. Aromatic saturation of distillates: an overview. Appl. Catal., A 1996, 137 (2), 203−223. (5) McVicker, G. B.; Daage, M.; Touvelle, M. S.; Hudson, C. W.; Klein, D. P.; Baird, W. C.; Cook, B. R.; Chen, J. G.; Hantzer, S.; Vaughan, D. E. W.; Ellis, E. S.; Feeley, O. C. Selective Ring Opening of Naphthenic Molecules. J. Catal. 2002, 210 (1), 137−148. (6) Gutiérrez, A.; Arandes, J. M.; Castaño, P.; Olazar, M.; Barona, A.; Bilbao, J. Effect of space velocity on the hydrocracking of Light Cycle Oil over a Pt−Pd/HY zeolite catalyst. Fuel Process. Technol. 2012, 95, 8−15. (7) Le Bihan, L.; Yoshimura, Y. Control of hydrodesulfurization and hydrodearomatization properties over bimetallic Pd−Pt catalysts supported on Yb-modifed USY zeolite. Fuel 2002, 81 (4), 491−494.

With the increasing of reaction temperature, naphthalene conversion was decreased while the yield of ring-opening products was increased. Because of kinetics and thermodynamics limitation, higher temperature decreases naphthalene hydrogenation reaction, but it can accelerate ring-opening reaction to form products. The NiMo/(Al2O3 + USY-F-AT) gives the highest yield of ring-opening products compared to NiMo/(Al2O3 + USY) both at 380 and 400 °C. For the hydrogenation of naphthalene, the conversion of naphthalene depends on its hydrogenation activity. In the hydrogenation reaction, the hydrogenation network of naphthalene is very complicated. The reaction network includes the sequential hydrogenation, ring-opening (ROP), isomerization, and dealkylation reactions. Hydrogenation reaction of naphthalene to produce tetralin is much easier than further hydrogenation of tetralin, so the yield of tetralin is higher than that of decalin.28 In addition, the acidity of the catalysts strongly affects the activity of ring-opening reaction.29 E

DOI: 10.1021/acs.iecr.9b00422 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research (8) Nagy, G.; Pölczmann, G.; Kalló, D.; Hancsók, J. Investigation of hydrodearomatization of gas oils on noble metal/support catalysts. Chem. Eng. J. 2009, 154 (1), 307−314. (9) Nagy, G.; Hancsók, J.; Varga, Z.; Pölczmann, G.; Kalló, D. Investigation of hydrodearomatization of prehydrogenated gas oil fractions on Pt−Pd/H-USY catalysts. Top. Catal. 2007, 45 (1), 195− 201. (10) Ma, H.; Yang, X.; Wen, G.; Tian, G.; Wang, L.; Xu, Y.; Wang, B.; Tian, Z.; Lin, L. Coupled hydrogenation and ring opening of tetralin on potassium modified Pt/USY catalysts. Catal. Lett. 2007, 116 (3), 149−154. (11) Ding, L.; Zheng, Y.; Zhang, Z.; Ring, Z.; Chen, J. Hydrotreating of light cycled oil using WNi/Al2O3 catalysts containing zeolite beta and/or chemically treated zeolite Y. J. Catal. 2006, 241 (2), 435−445. (12) Hédoire, C.; Louis, C.; Davidson, A.; Breysse, M.; Maugé, F.; Vrinat, M. Support effect in hydrotreating catalysts: hydrogenation properties of molybdenum sulfide supported on β-zeolites of various acidities. J. Catal. 2003, 220 (2), 433−441. (13) Corma, A.; Martínez, A.; Martínez-Soria, V. Catalytic Performance of the New Delaminated ITQ-2 Zeolite for Mild Hydrocracking and Aromatic Hydrogenation Processes. J. Catal. 2001, 200 (2), 259−269. (14) Contreras, R.; Ramírez, J.; Cuevas-García, R.; GutiérrezAlejandre, A.; Castillo-Villalón, P.; Macías, G.; Puente-Lee, I. Preparation and characterization of Pt/HMFI−SBA-15 hybrid catalyst for tetralin transformation. Catal. Today 2009, 148 (1), 49−54. (15) Arribas, M. A.; Concepción, P.; Martínez, A. The role of metal sites during the coupled hydrogenation and ring opening of tetralin on bifunctional Pt(Ir)/USY catalysts. Appl. Catal., A 2004, 267 (1), 111− 119. (16) Shimada, H.; Sato, K.; Honna, K.; Enomoto, T.; Ohshio, N. Design and development of Ti-modified zeolite-based catalyst for hydrocracking heavy petroleum. Catal. Today 2009, 141 (1), 43−51. (17) Yoshimura, Y.; Yasuda, H.; Sato, T.; Kijima, N.; Kameoka, T. Sulfur-tolerant Pd-Pt/Yb-USY zeolite catalysts used to reformulate diesel oils. Appl. Catal., A 2001, 207 (1), 303−307. (18) Maity, S. K.; Ancheyta, J. Carbon modified Y zeolite used as support material for hydroprocessing catalysts. Catal. Today 2010, 150 (3), 231−236. (19) Van Aelst, J.; Haouas, M.; Gobechiya, E.; Houthoofd, K.; Philippaerts, A.; Sree, S. P.; Kirschhock, C. E. A.; Jacobs, P.; Martens, J. A.; Sels, B. F.; Taulelle, F. Hierarchization of USY Zeolite by NH4OH. A Postsynthetic Process Investigated by NMR and XRD. J. Phys. Chem. C 2014, 118 (39), 22573−22582. (20) Ahn, J. H.; Kolvenbach, R.; Neudeck, C.; Al-Khattaf, S. S.; Jentys, A.; Lercher, J. A. Tailoring mesoscopically structured H-ZSM5 zeolites for toluene methylation. J. Catal. 2014, 311, 271−280. (21) van Laak, A. N. C.; Gosselink, R. W.; Sagala, S. L.; Meeldijk, J. D.; de Jongh, P. E.; de Jong, K. P. Alkaline treatment on commercially available aluminum rich mordenite. Appl. Catal., A 2010, 382 (1), 65−72. (22) Pérez-Ramírez, J.; Abelló, S.; Bonilla, A.; Groen, J. C. Tailored Mesoporosity Development in Zeolite Crystals by Partial Detemplation and Desilication. Adv. Funct. Mater. 2009, 19 (1), 164−172. (23) Zhao, L.; Shen, B.; Gao, J.; Xu, C. Investigation on the mechanism of diffusion in mesopore structured ZSM-5 and improved heavy oil conversion. J. Catal. 2008, 258 (1), 228−234. (24) Groen, J. C.; Jansen, J. C.; Moulijn, J. A.; Pérez-Ramírez, J. Optimal Aluminum-Assisted Mesoporosity Development in MFI Zeolites by Desilication. J. Phys. Chem. B 2004, 108 (35), 13062− 13065. (25) Yu, L.; Huang, S.; Miao, S.; Chen, F.; Zhang, S.; Liu, Z.; Xie, S.; Xu, L. A Facile Top-Down Protocol for Postsynthesis Modification of Hierarchical Aluminum-Rich MFI Zeolites. Chem. - Eur. J. 2015, 21 (3), 1048−1054. (26) Holm, M. S.; Svelle, S.; Joensen, F.; Beato, P.; Christensen, C. H.; Bordiga, S.; Bjørgen, M. Assessing the acid properties of desilicated ZSM-5 by FTIR using CO and 2,4,6-trimethylpyridine (collidine) as molecular probes. Appl. Catal., A 2009, 356 (1), 23−30.

(27) Ding, S.; Li, A.; Jiang, S.; Zhou, Y.; Wei, Q.; Zhou, W.; Huang, Y.; Yang, Q.; Fan, T. Niobium modification effects on hydrodesulfurization of 4,6-DMDBT catalyzed on Ni-Mo-S active sites: A combination of experiments and theoretical study. Fuel 2019, 237, 429−441. (28) Liu, Y.; Li, W.; Liu, C.; Cui, M.; Liu, C. Hydrogenation of Naphthalene on HY/MCM-41/γ-Al2O3 Supported Sulfided Ni-Mo-P Catalyst. Chin. J. Catal. 2004, 25 (7), 537−541. (29) Baeza, P.; Villarroel, M.; Á vila, P.; López Agudo, A.; Delmon, B.; Gil-Llambías, F. J. Spillover hydrogen mobility during Co−Mo catalyzed HDS in industrial-like conditions. Appl. Catal., A 2006, 304, 109−115.

F

DOI: 10.1021/acs.iecr.9b00422 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX