Fabrication of Metal–Organic Frameworks inside Silica Nanopores

Mar 14, 2018 - To date, the composites of MOFs with mesoporous materials like silica(37−39) and carbon nanotubes (CNTs)(40−42) have been reported...
0 downloads 9 Views 1MB Size
Subscriber access provided by Brought to you by ST ANDREWS UNIVERSITY LIBRARY

Surfaces, Interfaces, and Applications

Fabrication of Metal-Organic Frameworks inside Silica Nanopores with Significantly Enhanced Hydrostability and Catalytic Activity Jiahui Kou, and Lin-Bing Sun ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b01652 • Publication Date (Web): 14 Mar 2018 Downloaded from http://pubs.acs.org on March 15, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Jiahui Koua,b and Lin-Bing Sun*,a,c a

State Key Laboratory of Materials-Oriented Chemical Engineering, Jiangsu National Synergetic Innovation Center for Advanced Materials (SICAM), Nanjing Tech University, Nanjing 210009, China. b College of Materials Science and Engineering, Nanjing Tech University, Nanjing 210009, China. c College of Chemical Engineering, Nanjing Tech University, Nanjing 210009, China. ABSTRACT: Due to their diverse structure, high porosity, and tunable functionality, metal-organic frameworks (MOFs) are of great interest for diverse applications including catalysis. However, the poor hydrostability of most reported MOFs hinders their catalytic applications seriously. In addition, the development of an effective method to improve the catalytic activity of MOFs is another challenge. Here, we report for the first time the development of a double-solvent strategy to fabricate MOFs inside silica nanopores. A typical MOF (MOF-5) and a mesoporous silica with 2D hexagonal pore regularity (SBA-15) were first attempted. The double-solvent strategy is based on a hydrophobic solvent and a hydrophilic solution containing MOF precursors with a volume equal to or less than the pore volume of the support, so that the MOF can be formed selectively in the channels of support. Our results show that, upon confinement in silica nanopores, the hydrostability of MOF-5 is apparently improved. The framework of MOF-5 is destroyed obviously in a humid environment for 15 min, but that confined in SBA-15 is well preserved after 8 h. Moreover, the catalytic activity of the composite MOF-5@SBA-15 is superior to that of pure MOF-5 regarding activity and reaction rate. Under the catalysis of MOF-5@SBA-15, the conversion of benzyl bromide in the Friedel-Crafts alkylation reaction can reach 100% at 80 oC for 3 h, which is much higher than pure MOF-5 (61%) and SBA-15 (0%). We also demonstrate that the double-solvent strategy can be successfully extended to other MOFs, such as HKUST-1 and ZIF-8. Our work might open up an avenue for the improvement of stability and performance of MOFs which is highly expected for a variety of applications. Keywords: MOFs, mesoporous silica, hydrostability, confinement, double-solvent strategy ■INTRODUCTION Metal-organic frameworks (MOFs) are a class of crystalline materials construed by bridging organic ligands between their metallic moieties.1-6 As exciting materials with ultrahigh surface and diverse structure,7-9 MOFs are promising in a wide range of applications varied from gas separation10-13 and storage14-16 to drug delivery17,18 and catalysis,19-24 among others. However, many of them have been shown to be unstable upon exposure to moisture. It is known that the presence of metalligand coordination bonds makes MOFs more susceptible to hydrolysis than their inorganic analogues such as zeolites.25-29 Very tiny amount of moisture is able to attack the coordination bonds in MOFs, which results in the degradation of bonds as well as the damage of framework. MOF-5 is a typical example, its framework begins to degrade in 10 min upon exposure in humid environment.20 Therefore, the improvement of hydrostability is highly expected for the practical applications of MOFs. Since the discovery of MOFs, they have been employed as catalysts for diverse reactions because metal ions or clusters possessing coordinative vacancies can act as active sites. It is known that MOF crystals with a smaller size tend to show high catalytic activities, because of better exposure of active sites. However, preparation of small-sized MOFs is a challenging task, since traditional synthetic methods prefer to form large crystals. If the crystal size of MOFs can be reduced, it is expected to significantly improve the catalytic activities. As

for practical applications, it is extremely expected to create effective methods to enhance the hydrostability and catalytic activity of MOFs. Another group of porous materials that have attracted significant attention are mesoporous silicas.30,31 These silica materials with ordered nanopores in the range from 2 to 50 nm can be selected as supports for including a variety of guests inside the channels, such as metals, oxide nanoparticles, and drug macromolecules.32-36 Taking into account that MOFs have poor hydrostability, if they are incorporated into the nanopores of silica, the special microenvironment in nanopores is expected to enhance the hydrostability of MOFs. Moreover, the crystal size can be well controlled due to the confined space, which is of high potential for the formation of highly active catalysts. In addition, the mesoporous channels could make the active sites greatly accessible to the substrates and accelerate diffusion of both substrates and products. To date, the composites of MOFs with mesoporous materials like silica37-39 and carbon nanotubes (CNT)40-42 have been reported. The composites can be promising in various aspects like enhanced mechanical/chemical stability and novel gas adsorption behavior. Nevertheless, solvothermal synthesis is frequently used in literature, and this method is difficult (or not efficient) for the introduction of MOFs inside the nanopores of support. Here, we report the fabrication of MOFs in silica nanopores by employing a double-solvent strategy for the first time (Scheme 1). By use of the double-solvent strategy, the high

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

interfacial tension between hydrophobic solvent and hydrophilic solution (containing MOF precursors) drives the precursors to enter the hydrophilic nanopores. This leads to the construction of MOFs in the nanopores of SBA-15, while limits the formation of MOFs outside the pores. In contract, in the conventional single-solvent process, a large amount of solution containing MOF precursors is used, which results in the deposition of bulk MOFs on the outer surface of support. Our results show that both the symmetry of silica support and the structure of MOFs are well maintained in the composite. It is fascinating that the composite shows superior hydrostability in comparison with pure MOFs. After exposure to humid atmosphere, the confined MOFs well preserve their structure, which is quite different from bulk MOFs whose structure is seriously destroyed. Furthermore, MOFs confined in nanopores with reduced crystal sizes, along with accelerated mass transfer of silica nanopores, make them vastly active for Friedel-Crafts alkylation. We also demonstrate that the double-solvent strategy can be extended to fabrication of a series of MOFs (e.g. MOF-5, HKUST-1, and ZIF-8) in silica nanopores. Scheme 1. Fabrication of MOFs in Mesoporous Silica SBA-15 using the Double-Solvent Strategy. (a) Dispersion of SBA-15 in Hydrophobic n-Octane; (b) Addition of the Hydrophilic DMF Solution Containing MOF Precursors (Metal Ions and Ligands); (c) Formation of MOFs at 120 °C for 24 h.

■EXPERIMENTAL SECTION Chemicals. Zinc nitrate hexahydrate (Xilong, 99.5%), 1,4dicarboxybenzene (H2BDC, Aladdin, 99%), cupric nitrate trihydrate (Sinopharm, > 99.5%), 1,3,5-benzene-tricarboxylic acid (H3BTC, Sinopharm, > 98%), 2-methylimidazole (2MI, Adamas, 98%) N,N-dimethylformamide (DMF, Sinopharm, > 99.5%), n-octane (Lingfeng, > 95%), EO20PO70EO20(P123, Sigma-Aldrich, > 99%), tetraethylorthosilicate (TEOS, SigmaAldrich, 98%), hydrochloric acid (Lingfeng, > 99.5%), benzyl

Page 2 of 10

bromide (Sinopharm, > 97.5%), dichloromethane (Sinopharm, > 99.5%), and toluene (Lingfeng, > 99.5%) were utilized directly without further treatment. Deionized water was produced by a Milli-Q integral pure and ultrapure water system and utilized in all of the experiments. Materials Synthesis. MOF-5 was prepared by dissolving Zn(NO3)2·6H2O (0.78 g, 3mmol) and H2BDC (0.165 g, 1mmol) in DMF (10 mL). The mixture was transferred to a flask (50 mL) and heated to 120 °C. The temperature was kept at this temperature for 24 h. Then the flask was cooled to room temperature, the crystals located on the bottom of the flask were recovered and washed with the solvent DMF followed by immersing in CH2Cl2. CH2Cl2 was exchanged twice in the following two days. Lastly, the obtained crystals were stored in fresh CH2Cl2. Mesoporous silica SBA-15 was synthesized according to the reported method.43 Triblock copolymer Pluronic P123 (3.0 g) was dissolved in the mixed solution of 2 M HCl (90 g) and deionized water (22.5 g). After the solution was stirred for 1 h at room temperature, TEOS (6.4 g) was added dropwise to the solution with constant stirring at 40 °C and then the resulting solution was continuously stirred at 40 °C for 24 h. Afterwards, the milky reaction mixture was hydrothermally treated at 100 °C for 24 h. The resultant precipitate was filtered off, washed with water, and air-dried at room temperature. The composite MOF-5@SBA-15 was prepared via a double-solvent strategy. Briefly, Zn(NO3)2·6H2O (0.78 g, 3 mmol) and 1,4-BDC (0.165 g, 1 mmol) were dissolved in DMF (2 mL) as the precursor for MOF-5. Subsequently, SBA-15 (0.1 g) was suspended in dry n-octane (20 mL) as the hydrophobic solvent and the mixture was stirred for 30 min. The hydrophilic precursor solution (0.1 mL) was added dropwise over a period of 30 min with vigorous stirring. The resulting solution was continuously stirred at 120 °C for 24 h. After cooling to room temperature, the supernate was removed and the solids deposited on the bottom of flask were collected, washed with DMF, and immersed in fresh CH2Cl2. CH2Cl2 was exchanged twice during two days. Ultimately, the resulting material was stored in fresh CH2Cl2. The content of MOF-5 in the composite was detected by inductively coupled plasma (ICP), and the results show that the content of MOF-5 is 10.3wt%. Characterization. X-ray diffraction (XRD) analysis of samples was conducted using a Bruker D8 Avance diffractometer. Nitrogen adsorption isotherms were collected by use of ASAP2020 at 196 °C. The materials were degassed at 100 °C for about 4 h before measurement. The Brunauer-EmmettTeller (BET) surface areas of samples were estimated by utilizing the adsorption results with a relative pressure varied from 0.04 to 0.20. The pore size distributions were estimated by use of the Barrett-Joyner-Halenda (BJH) method based on the adsorption branches. The total pore volumes were obtained from the uptake at the pressure (p/p0) of around 0.99. A Nicolet Nexus 470 spectrometer was employed for the measurement of Fourier transform infrared (IR) spectra. The spectra were recorded by diluting the samples with KBr. Scanning electronic microscopy (SEM) and transmission electron microscope (TEM) were carried out on a Hitachi S4800 electron microscope. Thermogravimetric (TG) analysis was conducted on a thermobalance (STA-499C, NETZSCH), and the derivatives (DTG) can be obtained by software analysis. Typically, the materials were heated to 800 °C with a heating rate 10 °C·min−1 in nitrogen atmosphere. For ICP measurement, HCl was used to dissolve the materials. The contents of MOF-

ACS Paragon Plus Environment

Page 3 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

5 can be calculated from the amount of zinc in the materials. The hydrostability of samples was determined as follows. The materials were exposed to humid atmosphere with a relative humidity (RH) of 60% up to 8 h. Then the treated materials were detected by XRD. The hydrostability of materials was assessed by comparing the XRD patterns before and after treatment. Catalytic Reaction. The Friedel-Crafts alkylation reactions between benzyl bromide and toluene were run in a flask. Typically, benzyl bromide (1.5 mL, 12.4 mmol), toluene (4.0 mL, 37.3 mmol), and a prescribed amount of catalyst were added. In order to quantitatively compare the catalytic performance, 100 mg MOF-5@SBA-15 (containing 10 mg MOF-5), 10 mg MOF-5, or 90 mg SBA-15 was used for the reaction. The reaction was conducted at 80 °C for 3 h. After the reaction was finished, the reaction mixture was centrifuged. The obtained liquid was detected by gas chromatograph. The analysis was carried out by use of Agilent GC 7890A equipped with a flame ionization detector and a HP-5 column. ■ RESULTS

Structural Characterization. The first studied MOF is MOF-5 due to its poor hydrostability. And a typical mesoporous silica, SBA-15, was employed as the host. Figure 1a shows the low-angle XRD patterns of MOF-5, MOF-5@SBA15, and SBA-15. There are no peaks in the pattern of MOF-5. The pattern of SBA-15 shows three well-resolved peaks which can be indexed as (100), (110) and (200) diffraction peaks associated with p6mm hexagonal symmetry.44-47 Although the diffraction intensity of MOF-5@SBA-15 decreases a little which is due to the presence of MOF-5 crystals in the SBA-15 channels, the mesoporous hexagonal structure is obviously remained. The wide-angle XRD patterns of the materials are shown in Figure 1b. MOF-5 exhibits well-defined diffraction lines, which is in line with the reported results.48-50 For SBAo 15, there is a broad peak at about 23 , which is caused by its amorphous walls. After incorporating MOF-5 crystals into the channels of SBA-15, the pattern has the characteristics of both MOF-5 and SBA-15. This indicates that the crystalline framework of MOF-5 is formed successfully in the channels of SBA-15. Furthermore, the intensity of the XRD peaks is weak in the composite, which indicates the small dimension of MOF-5 in the composite. Further information on the individual MOF-5 and SBA-15 as well as the composite MOF-5@SBA-15 is given by IR spectra (Figure 2). For MOF-5, the asymmetric stretching of carboxylate groups (COO) of BDC appear in around 1625 and 1500 cm−1 and symmetric stretching appear around 1389 cm−1.51,52 The typical vibration mode of SBA-15 is observed. The OH vibration of the adsorbed water appears at 1630 cm−1, while the band between 1300-1000 cm−1 is due to the SiOSi bond. The band observed at 962 cm−1 is attributed to SiOH, whereas the band at 462 cm−1 can be ascribed to SiO vibration. The IR spectrum of the composite has the characteristics of MOF-5 and SBA-15, indicating the successful formation of MOF-5 in SBA-15.

Figure 1. (a) Low-angle and (b) wide-angle XRD patterns of MOF-5, MOF-5@SBA-15, and SBA-15.

Figure 2. IR spectra of MOF-5, MOF-5@SBA-15, and SBA-15.

Figure 3a gives N2 adsorption-desorption isotherms of MOF-5 and SBA-15 as well as MOF-5@SBA-15. The isotherm for MOF-5 is of type I, indicative of the microporous feature. The SBA-15 shows type IV with an H1 hysteresis loop, which is the typical characteristic of a material with cylindrical mesopores.53,54 For MOF-5@SBA-15, the isotherm shows the characteristic of SBA-15. In Figure 3b, a pore size distribution at 8 nm is observed for SBA-15. Due to partial filling of the mesopores, there is a decrease in the pore size of MOF-5@SBA-15. Further textual parameters are listed in Table S1. Interesting results can be found from microporous

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and mesoporous volumes. The composite MOF-5@SBA-15 has a mesopore volume of 0.68 cm3·g1, while the mesoporous volume of MOF-5 is negligible. In addition, the composite MOF-5@SBA-15 shows a microporous volume of 0.20 cm3·g1, which is higher than that of SBA-15 (0.14 cm3·g1). For the ratio of Vmicro/Vtot, the composite MOF-5@SBA-15 presents a value of 0.23, which is obviously larger than that of SBA-15 (0.12). These results suggest that MOF-5 is successfully incorporated into SBA-15; and in comparison with pure SBA-15 more micropores are generated due to the incorporation of microporous material MOF-5. Additionally, abundant mesopores in the composite MOF-5@SBA-15 is beneficial to mass transfer as discussed below.

Page 4 of 10

nels of SBA-15. The size of these particles are much smaller than bulk MOF-5 synthesized by the traditional method. Elemental mappings were also taken and shown in Figure 5. In addition to Si and O derived from SBA-15, the element Zn stemmed from MOF-5 is detected. This further proved that MOF-5 is successfully grown in mesoporous silica SBA-15.

Figure 4. SEM and TEM images of (a, b) MOF-5, (c, d) MOF5@SBA-15, and (e, f) SBA-15.

Figure 3. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of MOF-5, MOF-5@SBA-15, and SBA-15.

Figure 5. STEM image of the composite MOF-5@SBA-15 and elemental mapping images of Si, O, and Zn. The domain in the red box was used for elemental analysis, and the domain in the yellow box was used for correction.

The morphology and pore structure of materials were studied using SEM and TEM as shown in Figure 4. MOF-5 presents cubic-like morphology and the crystal width is about 10 μm (Figure 4a). SBA-15 exhibits the typical wheat-like morphology and consists of aggregates of rope-like particles with relatively uniform size (Figure 4e). The composite MOF5@SBA-15 shows a similar morphology to SBA-15 (Figure 4c). No MOF particles scattered on the outer surface are observed. In the TEM images, the crystal of MOF-5 is damaged (Figure 4b), because it is unstable during irradiation with high energy X-ray. SBA-15 displays well-ordered hexagonal pore channels (Figure 4f). After the introduction of MOF-5, the cylindrical shape of pores and the nearly perfect 2D hexagonal order are maintained (Figure 4d). It is worth mentioning that the MOF-5 particles (dark spots) are visible inside the chan-

Through the aforementioned results, it is clear that the double-solvent strategy is effective in incorporating MOF-5 into silica nanopores. After the incorporation of MOF-5, the pore regularity of SBA-15 is well-preserved. The newly formed composite MOF-5@SBA-15 thus exhibits the characteristic of microporous and mesoporous materials. The MOF crystals confined in nanopores show interesting stability and catalytic performance as demonstrated as follows. Stability Examination. TG technique was employed to examine the thermal stability of samples. MOF-5 shows two weight loss steps (Figure 6a,). The first weight loss less than 200 °C is caused by solvent removal. The second main weight loss is due to the degradation of organic ligands, which corresponds to an obvious DTG peak at 530 °C (Figure 6b). It

ACS Paragon Plus Environment

Page 5 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

means that MOF-5 has good thermal stability. For SBA-15, the initial weight loss centered at about 100 °C is caused by the removal of physically adsorbed water. The second step between 150 and 300 °C is due to the decomposition of partial organic groups and removal of hydration water. There is a slight weight loss between 300 and 600 °C which corresponds to the combustion of residual carbon. For the composite, the weight loss trend is similar to MOF-5. However, the predominant weight loss takes place at 480 °C, which is lower than bulk MOF-5. The reason might be that the size of MOF crystals incorporated into mesoporous supports is much smaller than the size of the bulk MOF-5.

Figure 6. (a) TG and (b) DTG curves of MOF-5, MOF-5@SBA-15, and SBA-15.

The hydrostability is quite essential from the viewpoint of practical applications. In order to exam the hydrostability, the samples were put in a humid environment (RH = 60%) at room temperature for different time. XRD patterns of related samples after treatment are employed to determine the hydrostability. Figure 7a shows that significant loss of the diffraco tion peaks (e.g. the peak at 9.7 ) and the emergence of new o diffraction peaks (e.g. the peak at 8.9 ) can be observed for treated MOF-5. In fact, upon exposure to humid atmosphere MOF-5 is converted to MOF-69c and related derivatives.55,56 In the presence of moisture, MOF-69c can be transformed to another structure ZnBDC∙xH2O, where x is between 1 and 2. The XRD pattern of ZnBDC∙xH2O is reported in literature, but the exact crystal structure is not obtained until now. 56 This is different from the destruction of crystalline structure that often gives amorphous characteristic. The emergence of new charac-

teristic peaks of ZnBDC∙xH2O, together with the loss of structural peaks of MOF-5, can thus reflect the poor hydrostability of MOF-5.57,58 Hence, the variation in the XRD patterns demonstrates the destruction of crystalline structure. Surprisingly, the powder XRD pattern of MOF-5@SBA-15 remains unchanged, and the original peaks keep constant and no new peaks emerge (Figure 7b). This implies the enhanced hydrostability of the composite MOF-5@SBA-15. Apparently, the hydrostability of MOF-5 is greatly improved upon confinement in the mesopores of SBA-15.

Figure 7. Wide-angle XRD patterns of (a) MOF-5 and (b) MOF5@SBA-15 after being exposed to humid atmosphere (RH = 60%).

In short, the decomposition temperature of MOF-5 in the composite is somewhat lower than bulk MOF-5 due to the small-sized MOF crystals. The hydrostability of MOF-5 improves obviously upon confinement and the MOF is stable under humid atmosphere up to 8 h while bulk MOF-5 collapses within 15 min. Catalytic Performance. In chemical industries, FriedelCrafts alkylation of aromatic compounds is an important reaction.59,60 These liquid-phase reactions have been traditionally catalyzed by Lewis acids such as AlCl3, ZnCl2, and FeCl3.61-65 However, these catalysts present undesirable economic and environmental issues. Therefore, solid acid catalysts have been investigated for the Friedel-Crafts alkylation. Since MOF-5 has high surface area as well as coordination unsaturated open metal sites, it was used to catalyze Friedel-Crafts alkylation. The obtained materials were assessed for its catalytic activity in the Friedel-Crafts alkylation by studying reaction of benzyl

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

bromide with toluene. As shown in Figure 8, no benzyl bromide was converted at all even after 3 h on the silica support SBA-15 because there is no active sites. On MOF-5, the conversion of benzyl bromide is 61% at 3 h, indicating that there exists active sites for the reaction. The incorporation of MOF5 into nanopores adjusts the activity apparently. Under the catalysis of MOF-5@SBA-15, the conversion of benzyl bromide reaches 100% under the same reaction conditions, which is much higher than that over MOF-5. Remarkably, the reaction rate is rather different in the case of MOF-5 before and after being incorporated into SBA-15. Under the catalysis of MOF-5, the conversion of benzyl bromide is 36% at 1.5 h, which accounts for 59% of the conversion at 3 h (61%). Nevertheless, MOF-5@SBA-15 can convert 80% of benzyl bromide at 1.5 h, which is 80% of the conversion at 3 h (100%). The recyclability of MOF@SBA-15 was studied as well. After three recycle, the conversion of benzyl bromide is 85% at 3 h. This activity is kind of lower than that over the fresh catalyst, while most of catalytic activity is remained. Based on these results, it is safe to say that the incorporation of MOF-5 into SBA-15 leads to the increase of reaction rate. In short, the catalytic performance of MOF-5 is significantly enhanced upon confinement in silica nanopores regarding catalytic activity as well as reaction rate.

Page 6 of 10

the results of pore size distributions indicate that the introduction of MOF-5 leads to the decrease of pore size distributions (Figure 3b). If MOF-5 is located outside the pores, the pore size of SBA-15 would keep constant. Second, the elemental mapping images show that Zn derived from MOF-5 is well distributed in the whole sample. Third, both thermal stability and hydrostability change obviously for MOF-5 confined in SBA-15 (Figures 6 and 7). If the MOF-5 crystals are outside the pores, their stability should not be changed too much, if any. It is noticeable that the incorporation of MOF-5 into SBA-15 improves the hydrostability significantly. The structure of bulk MOF-5 is damaged seriously after being exposed to a humid atmosphere within 15 min, whereas MOF-5 confined in mesopores is stable up to 8 h. The main reason may be that the silica walls of SBA-15 could provide protection for MOF-5, which prevents the MOF crystals confined in mesopores from being attacked by the water vapor in the atmosphere. Consequently, the composite MOF-5@SBA-15 shows enhanced hydrostability. Our results prove that the composite is highly active in Friedel-Crafts alkylation. The conversion over the composite (100% at 3 h) is obviously higher than that over MOF-5 (only 61% at 3 h). It is well-known that the catalytic activity generally increases with the decrease of catalyst size, as smaller MOFs have higher surface areas available for reactants. In our study, we have proved that the size of MOF-5 crystals grown in silica mesopores is much smaller than the bulk ones. Thus the reactant molecules are more readily accessible to the catalyst active sites. In addition, the mesoporous structure can accelerate the diffusion of reactant/ product molecules. These properties make the mesoporous silica-confined MOFs highly promising in various applications.

Figure 8. Conversion in Friedel-Crafts alkylation of benzyl bromide with toluene over MOF-5, MOF-5@SBA-15, and SBA-15. ■ DISCUSSION

Due to their diverse structure, high surface area, and adjustable functionality, MOFs are highly potential in diverse applications. Nevertheless, there are few reports concerning the practical applications of MOFs, because of their poor hydrostability. To overcome this shortcoming, we report a doublesolvent strategy to fabricate MOF-5 within the mesopores of SBA-15, a typical mesoporous silica. The double-solvent strategy is based on a hydrophilic solution and a hydrophobic solvent. The former contains the precursor with a volume set equal to or less than the pore volume of the support, which can be absorbed within the hydrophilic support pores; and the latter, in a large amount, plays an important role to suspend the support and facilitate the synthetic process. This leads to the construction of dispersed MOFs in the SBA-15 channels, avoids the deposition of MOFs outside the pores, and produces MOFs with small size. There are some evidences that can prove the introduction of MOF-5 to the pores of SBA-15. First,

Figure 9. (a) Wide-angle XRD patterns of HKUST-1, HKUST1@SBA-15, and SBA-15. Photographs for the fabrication of HKUST1 in mesoporous silica SBA-15 using the double-solvent strategy: (b) Dispersion of SBA-15 in hydrophobic n-octane; (c) addition of the hydrophilic DMF solution containing HKUST-1 precursors (metal ions and ligands); (d) formation of MOFs at 120 °C for 24 h; (e) washing twice with DMF and methanol.

ACS Paragon Plus Environment

Page 7 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

To generalize the incorporation of MOFs into silica nanopores, the fabrication of other MOFs by use of the doublesolvent strategy was attempted. In addition to MOF-5, two other typical MOFs (i.e. HKUST-1 and ZIF-8) were examined. The synthetic process for HKUST-1@SBA-15 was recorded as shown in Figure 9. After dropwise adding the blue hydrophilic solution of MOF precursors, the white mesoporous silica becomes light blue, while the supernate is colorless (Figure 9c). This phenomenon clearly proves that HKUST-1 precursor has been completely adsorbed in the pores of SBA-15. The results of ICP show that about 11.2 wt% of HKUST-1 is introduced to mesoporous silica. Figure 9d shows the formation of composite after reacting at 120 °C for 24 h. As can be seen from the picture, the color of the sample is deepened, indicating the successful formation of MOF crystal. After washing, the composite was stored in CH2Cl2 (Figure 9e). Moreover, the wide-angle XRD patterns (Figure 9a), low-angle XRD patterns (Figure S1), and IR spectra (Figure S2) attest the successful synthesis of HKUST-1@SBA-15. Further information is given in TG and DTG curves (Figure S3), the decomposition temperature of HKUST-1@SBA-15 is slightly lower than that of HKUST-1, which agrees with the results reported in MOF5@SBA-15. The decreased thermal stability actually indicates the smaller size of crystals in comparison with the bulk ones. By utilizing the double-solvent strategy, the composite ZIF8@SBA-15 is also synthesized successfully according to the low-angle XRD patterns (Figure S4), wide-angle XRD patterns (Figure S5), IR spectra (Figure S6), and TG and DTG curves (Figure S7). The abovementioned results thus demonstrate that the double-solvent strategy can be extended to the incorporation of different MOFs into mesoporous silica. ■ CONCLUSIONS A facile and efficient double-solvent strategy has been developed for the construction of MOFs inside the silica nanopores, for the first time. In comparison with bulk MOFs, MOFs confined in silica nanopores have greatly improved hydrostability. Because of the nanometer sized MOF crystals in nanopores, the active sites in MOFs become easily accessible. Also, the mesoporous channels of support favor the mass transfer during the reactions. As a result, the composite exhibits excellent catalytic performance in Friedel-Crafts alkylation reactions regarding catalytic activity and reaction rate, which is much superior to bulk MOFs. We also demonstrate that the doublesolvent strategy can be extended to the fabrication of a series of MOFs inside the nanopores of silica support. The present study may bring light to the construction of functional materials combining MOFs and mesoporous materials, resulting in the composites with enhanced stability and activity for a variety of applications. ■ ASSOCIATED CONTENT Textual parameters, XRD patterns, IR spectra, and TG curves of different samples. This material is available free of charge via the Internet at http://pubs.acs.org.

■ AUTHOR INFORMATION Corresponding Author * E-mail: [email protected] (L.-B. S.)

■ ACKNOWLEDGMENTS

We acknowledge financial support of this work by the National Natural Science Foundation of China (21722606 and 21576137).

■ REFERENCES (1) Cook, T. R.; Zheng, Y. R.; Stang, P. J. Metal-Organic Frameworks and Self-Assembled Supramolecular Coordination Complexes: Comparing and Contrasting the Design, Synthesis, and Functionality of Metal-Organic Materials. Chem. Rev. 2013, 113, 734-777. (2) Schneemann, A.; Bloch, E. D.; Henke, S.; Llewellyn, P. L.; Long, J. R.; Fischer, R. A. Influence of Solvent-Like Sidechains on the Adsorption of Light Hydrocarbons in MetalOrganic Frameworks. Chem.-Eur. J. 2015, 21, 18764-18769. (3) Chakrabarty, R.; Mukherjee, P. S.; Stang, P. J. Supramolecular Coordination: Self-Assembly of Finite Twoand Three-Dimensional Ensembles. Chem. Rev. 2011, 111, 6810-6918. (4) Schoedel, A.; Ji, Z.; Yaghi, O. M. The Role of Metal– Organic Frameworks in a Carbon-Neutral Energy Cycle. Nature Energy 2016, 1, 16034-16047. (5) Sun, L. B.; Li, J. R.; Park, J.; Zhou, H. C. Cooperative Template-Directed Assembly of Mesoporous Metal-Organic Frameworks. J. Am. Chem. Soc. 2012, 134, 126-129. (6) Furukawa, H.; Cordova, K. E.; O'Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 974-986. (7) Qiu, X.; Zhong, W.; Bai, C.; Li, Y. Encapsulation of a Metal-Organic Polyhedral in the Pores of a Metal-Organic Framework. J. Am. Chem. Soc. 2016, 138, 1138-1141. (8) Andrew Lin, K.-Y.; Chang, H.-A.; Chen, B.-J. MultiFunctional MOF-Derived Magnetic Carbon Sponge. J. Mater. Chem. A 2016, 4, 13611-13625. (9) Sun, Q.; He, H.; Gao, W.-Y.; Aguila, B.; Wojtas, L.; Dai, Z.; Li, J.; Chen, Y.-S.; Xiao, F.-S.; Ma, S. Imparting Amphiphobicity on Single-Crystalline Porous Materials. Nat. Commun. 2016, 7, 13300-13307. (10) Wang, X.; Chi, C.; Zhang, K.; Qian, Y.; Gupta, K. M.; Kang, Z.; Jiang, J.; Zhao, D. Reversed Thermo-Switchable Molecular Sieving Membranes Composed of TwoDimensional Metal-Organic Nanosheets for Gas Separation. Nat. Commun. 2017, 8, 14460-14470. (11) Chen, C.; Li, B.; Zhou, L.; Xia, Z.; Feng, N.; Ding, J.; Wang, L.; Wan, H.; Guan, G. Synthesis of Hierarchically Structured Hybrid Materials by Controlled Self-Assembly of Metal–Organic Framework with Mesoporous Silica for CO2 Adsorption. ACS Appl. Mater. Interf. 2017, 9, 23060-23071. (12) Mazaj, M.; Cendak, T.; Buscarino, G.; Todaro, M.; Zabukovec Logar, N. Confined Crystallization of a HKUST-1 Metal-OrganicFramework within Mesostructured Silica with Enhanced Structural Resistance towards Water. J. Mater. Chem. A 2017, 5, 22305-22315. (13) Lee, K.; Isley, W. C.; Dzubak, A. L.; Verma, P.; Stoneburner, S. J.; Lin, L. C.; Howe, J. D.; Bloch, E. D.; Reed, D. A.; Hudson, M. R.; Brown, C. M.; Long, J. R.; Neaton, J. B.; Smit, B.; Cramer, C. J.; Truhlar, D. G.; Gagliardi, L. Design of a Metal-Organic Framework with Enhanced Back Bonding for Separation of N2 and CH4. J. Am. Chem. Soc. 2014, 136, 698-704. (14) Mason, J. A.; Veenstra, M.; Long, J. R. Evaluating Metal-Organic Frameworks for Natural Gas Storage. Chem. Sci. 2014, 5, 32-51. (15) Yang, J.; Sudik, A.; Wolverton, C.; Siegel, D. J. High Capacity Hydrogen Storage Materials: Attributes for

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Automotive Applications and Techniques for Materials Discovery. Chem. Soc. Rev. 2010, 39, 656-675. (16) Yan, Y.; Yang, S. H.; Blake, A. J.; Schroder, M. Studies on Metal-Organic Frameworks of Cu(ii) with Isophthalate Linkers for Hydrogen Storage. Acc. Chem. Res. 2014, 47, 296307. (17) Taylor Pashow; L., K. M.; Della Rocca, J.; Xie, Z.; Tran, S.; Lin, W. Postsynthetic Modifications of IronCarboxylate Nanoscale Metal-Organic Frameworks for Imaging and Drug Delivery. J. Am. Chem. Soc. 2009, 131, 14261-14263. (18) Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J. F.; Heurtaux, D.; Clayette, P.; Kreuz, C.; Chang, J. S.; Hwang, Y. K.; Marsaud, V.; Bories, P.-N.; Cynober, L.; Gil, S.; Ferey, G.; Couvreur, P.; Gref, R. Porous Metal-Organic-Framework Nanoscale Carriers as a Potential Platform for Drug Delivery and Imaging. Nat. Mater. 2010, 9, 172-178. (19) Ahmed, I.; Khan, N. A.; Jhung, S. H. Graphite Oxide/Metal–Organic Framework (MIL-101): Remarkable Performance in the Adsorptive Denitrogenation of Model Fuels. Inorg. Chem. 2013, 52, 14155-14161. (20) Hansen, R. E.; Das, S. Biomimetic Di-Manganese Catalyst Cage-Isolated in a MOF: Robust Catalyst for Water Oxidation with Ce(IV), A Non-O-Donating Oxidant. Energy Environ. Sci. 2014, 7, 317-322. (21) Mo, K.; Yang, Y. H.; Cui, Y. A Homochiral MetalOrganic Framework as an Effective Asymmetric Catalyst for Cyanohydrin Synthesis. J. Am. Chem. Soc. 2014, 136, 17461749. (22) Zhao, M.; Ou, S.; Wu, C. D. Porous Metal-Organic Frameworks for Heterogeneous Biomimetic Catalysis. Acc. Chem. Res. 2014, 47, 1199-1207. (23) Zhu, L.; Liu, X.-Q.; Jiang, H.-L.; Sun, L.-B. Metal– Organic Frameworks for Heterogeneous Basic Catalysis. Chem. Rev. 2017, 117, 8129-8176. (24) Sun, L.-B.; Liu, X.-Q.; Zhou, H.-C. Design and Fabrication of Mesoporous Heterogeneous Basic Catalysts. Chem. Soc. Rev. 2015, 44, 5092-5147. (25) Burtch, N. C.; Jasuja, H.; Walton, K. S. Water Stability and Adsorption in Metal–Organic Frameworks. Chem. Rev. 2014, 114, 10575-10612. (26) Howarth, A. J.; Liu, Y.; Li, P.; Li, Z.; Wang, T. C.; Hupp, J. T.; Farha, O. K. Chemical, Thermal and Mechanical Stabilities of Metal–Organic Frameworks. Nat. Rev. Mater. 2016, 1, 15018-15032. (27) He, H.; Sun, Q.; Gao, W.; Perman, J. A.; Sun, F.; Zhu, G.; Aguila, B.; Forrest, K.; Space, B.; Ma, S. A Stable MetalOrganic Framework Featuring Local Buffer Environment for Carbon Dioxide Fixation. Angew. Chem. Int. Ed. 2018, DOI: 10.1002/anie.201801122. (28) Gao, W.-Y.; Cai, R.; Pham, T.; Forrest, K. A.; Hogan, A.; Nugent, P.; Williams, K.; Wojtas, L.; Luebke, R.; Weseliński, Ł. J.; Zaworotko, M. J.; Space, B.; Chen, Y.-S.; Eddaoudi, M.; Shi, X.; Ma, S. Remote Stabilization of Copper Paddlewheel Based Molecular Building Blocks in Metal– Organic Frameworks. Chem. Mater. 2015, 27, 2144-2151. (29) Sun, Y.; Sun, Q.; Huang, H.; Aguila, B.; Niu, Z.; Perman, J. A.; Ma, S. A Molecular-Level Superhydrophobic External Surface to Improve the Stability of Metal-Organic Frameworks. J. Mater. Chem. A 2017, 5, 18770-18776. (30) Jiang, W.-J.; Yin, Y.; Liu, X.-Q.; Yin, X.-Q.; Shi, Y.-Q.; Sun, L.-B. Fabrication of Supported Cuprous Sites at Low

Page 8 of 10

Temperatures: An efficient, Controllable Strategy Using Vapor-Induced Reduction. J. Am. Chem. Soc. 2013, 135, 8137-8140. (31) Lakhi, K. S.; Park, D.-H.; Al-Bahily, K.; Cha, W.; Viswanathan, B.; Choy, J.-H.; Vinu, A. Mesoporous Carbon Nitrides: Synthesis, Functionalization, and Applications. Chem. Soc. Rev. 2017, 46, 72-101. (32) Schlossbauer, A.; Schaffert, D.; Kecht, J.; Wagner, E.; Bein, T. Click Chemistry for High-Density Biofunctionalization of Mesoporous Silica. J. Am. Chem. Soc. 2008, 130, 12558-12559. (33) Qi, H.; Shopsowitz, K. E.; Hamad, W. Y.; MacLachlan, M. J. Chiral Nematic Assemblies of Silver Nanoparticles in Mesoporous Silica Thin Films. J. Am. Chem. Soc. 2011, 133, 3728-3731. (34) Park, C.; Kim, H.; Kim, S.; Kim, C. Enzyme Responsive Nanocontainers with Cyclodextrin Gatekeepers and Synergistic Effects in Release of Guests. J. Am. Chem. Soc. 2009, 131, 16614-16615. (35) Coll, C.; Mondragon, L.; Martinez Manez, R.; Sancenon, F.; Dolores Marcos, M.; Soto, J.; Amoros, P.; Perez Paya, E. Enzyme-Mediated Controlled Release Systems by Anchoring Peptide Sequences on Mesoporous Silica Supports. Angew. Chem. Int. Ed. 2011, 50, 2138-2140. (36) Climent, E.; Martinez Manez, R.; Sancenon, F.; Marcos, M. D.; Soto, J.; Maquieira, A.; Amoros, P. Controlled Delivery using Oligonucleotide-Capped Mesoporous Silica Nanoparticles. Angew. Chem. Int. Ed. 2010, 49, 7281-7283. (37) Ahmed, A.; Forster, M.; Clowes, R.; Bradshaw, D.; Myers, P.; Zhang, H. F. Silica SOS@HKUST-1 Composite Microspheres as Easily Packed Stationary Phases for Fast Separation. J. Mater. Chem. A 2013, 1, 3276-3286. (38) Zhao, M. T.; Deng, K.; He, L. C.; Liu, Y.; Li, G. D.; Zhao, H. J.; Tang, Z. Y. Core-Shell Palladium Nanoparticle@Metal-Organic Frameworks as Multifunctional Catalysts for Cascade Reactions. J. Am. Chem. Soc. 2014, 136, 1738-1741. (39) Zhu, Q. L.; Xu, Q. Metal-Organic Framework Composites. Chem. Soc. Rev. 2014, 43, 5468-5512. (40) Lin, R. J.; Ge, L.; Liu, S. M.; Rudolph, V.; Zhu, Z. H. Mixed-Matrix Membranes with Metal-Organic FrameworkDecorated CNT Fillers for Efficient CO2 Separation. ACS Appl. Mater. Interf. 2015, 7, 14750-14757. (41) Wen, P.; Gong, P. W.; Sun, J. F.; Wang, J. Q.; Yang, S. R. Design and Synthesis of Ni-MOF/CNT Composites and rGO/Carbon Nitride Composites for an Asymmetric Supercapacitor with High Energy and Power Density. J. Mater. Chem. A 2015, 3, 13874-13883. (42) Han, T. T.; Xiao, Y. L.; Tong, M. M.; Huang, H. L.; Liu, D. H.; Wang, L. Y.; Zhong, C. L. Synthesis of CNT@MIL68(Al) Composites with Improved Adsorption Capacity for Phenol in Aqueous Solution. Chem. Eng. J. 2015, 275, 134141. (43) Zhao, D. Y.; Feng, J. L.; Huo, Q. S.; Melosh, N.; Fredrickson, G. H.; Chmelka, B. F.; Stucky, G. D. Triblock Copolymer Syntheses of Mesoporous Silica with Periodic 50 to 300 Angstrom Pores. Science 1998, 279, 548-552. (44) Gu, D.; Li, W.; Wang, F.; Bongard, H.; Spliethoff, B.; Schmidt, W.; Weidenthaler, C.; Xia, Y.; Zhao, D.; Schueth, F. Controllable Synthesis of Mesoporous Peapod-Like Co3O4@Carbon Nanotube Arrays for High-Performance Lithium-Ion Batteries. Angew. Chem. Int. Ed. 2015, 54, 70607064.

ACS Paragon Plus Environment

Page 9 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(45) Li, T. T.; Sun, L. B.; Liu, X. Y.; Sun, Y. H.; Song, X. L.; Liu, X. Q. Isolated Lithium Sites Supported on Mesoporous Silica: A Novel Solid Strong Base with High Catalytic Activity. Chem. Commun. 2012, 48, 6423-6425. (46) Sun, Y. H.; Sun, L. B.; Li, T. T.; Liu, X. Q. Modulating the Host Nature by Coating Alumina: A Strategy to Promote Potassium Nitrate Decomposition and Superbasicity Generation on Mesoporous Silica SBA-15. J. Phys. Chem. C 2010, 114, 18988-18995. (47) Yin, Y.; Tan, P.; Liu, X.-Q.; Zhu, J.; Sun, L.-B. Constructing a Confined Space in Silica Nanopores: An Ideal Platform for the Formation and Dispersion of Cuprous Sites. J. Mater. Chem. A 2014, 2, 3399-3406. (48) Ming, Y.; Purewal, J.; Yang, J.; Xu, C.; Soltis, R.; Warner, J.; Veenstra, M.; Gaab, M.; Mueller, U.; Siegel, D. J. Kinetic Stability of MOF-5 in Humid Environments: Impact of Powder Densification, Humidity Level, and Exposure Time. Langmuir 2015, 31, 4988-4995. (49) Zhang, S.-Y.; Li, D.; Guo, D.; Zhang, H.; Shi, W.; Cheng, P.; Wojtas, L.; Zaworotko, M. J. Synthesis of a Chiral Crystal Form of MOF-5, CMOF-5, by Chiral Induction. J. Am. Chem. Soc. 2015, 137, 15406-15409. (50) Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing Catalytic Activity and Stability for CO2 Methanation on Ni@MOF-5 via Control of Active Species Dispersion. Chem. Commun. 2015, 51, 1728-1731. (51) Lu, L.; Li, X. Y.; Liu, X. Q.; Wang, Z. M.; Sun, L. B. Enhancing the Hydrostability and Catalytic Performance of Metal–Organic Frameworks by Hybridizing with Attapulgite, A Natural Clay. J. Mater. Chem. A 2015, 3, 6998-7005. (52) Sabouni, R.; Kazemian, H.; Rohani, S. A Novel Combined Manufacturing Technique for Rapid Production of IrMOF-1 Using Ultrasound and Microwave Energies. Chem. Eng. J. 2010, 165, 966-973. (53) Reichhardt, N.; Kjellman, T.; Sakeye, M.; Paulsen, F.; Smatt, J. H.; Linden, M.; Alfredsson, V. Removal of Intrawall Pores in SBA-15 by Selective Modification. Chem. Mater. 2011, 23, 3400-3403. (54) Tian, W. H.; Sun, L. B.; Song, X. L.; Liu, X. Q.; Yin, Y.; He, G. S. Adsorptive Desulfurization by Copper Species within Confined Space. Langmuir 2010, 26, 17398-17404. (55) Zhang, M.; Bosch, M.; Gentle, T.; Zhou, H.-C. Rational Design of Metal-Organic Frameworks with Anticipated

Porosities and Functionalities. Crystengcomm 2014, 16, 40694083. (56) Hausdorf, S.; Wagler, J.; Mossig, R.; Mertens, F. O. R. L. Proton and Water Activity-Controlled Structure Formation in Zinc Carboxylate-Based Metal Organic Frameworks. J. Phys. Chem. A 2008, 112, 7567-7576. (57) Wu, C.-M.; Rathi, M.; Ahrenkiel, S. P.; Koodali, R. T.; Wang, Z. Facile Synthesis of MOF-5 Confined in SBA-15 Hybrid Material with Enhanced Hydrostability. Chem. Commun. 2013, 49, 1223-1225. (58) Kaye, S. S.; Dailly, A.; Yaghi, O. M.; Long, J. R. Impact of Preparation and Handling on the Hydrogen Storage Properties of Zn4O(1,4-Benzenedicarboxylate)3 (MOF-5). J. Am. Chem. Soc. 2007, 129, 14176-14177. (59) Phan, N. T. S.; Le, K. K. A.; Phan, T. D. MOF-5 as an Efficient Heterogeneous Catalyst for Friedel-Crafts Alkylation Reactions. Appl. Catal., A 2010, 382, 246-253. (60) Shimizu, K.; Niimi, K.; Satsuma, A. Polyvalent-Metal Salts of Heteropolyacid as Catalyst for Friedel-Crafts Alkylation Reactions. Appl. Catal., A 2008, 349, 1-5. (61) Bachari, K.; Millet, J. M. M.; Benaichouba, B.; Cherifi, O.; Figueras, F. Benzylation of Benzene by Benzyl Chloride over Iron Mesoporous Molecular Sieves Materials. J. Catal. 2004, 221, 55-61. (62) Hentit, H.; Bachari, K.; Ouali, M. S.; Womes, M.; Benaichouba, B.; Jumas, J. C. Alkylation of Benzene and Other Aromatics by Benzyl Chloride over Iron-Containing Aluminophosphate Molecular Sieves. J. Mol. Catal. A-Chem. 2007, 275, 158-166. (63) Lee, M.; Seo, Y.; Shin, H. S.; Jo, B.; Ryoo, R. Anatase TiO2 Nanosheets with Surface Acid Sites for Friedel-Crafts Alkylation. Microporous Mesoporous Mater. 2016, 222, 185191. (64) Liang, X.; Jiang, S. Z.; Wei, K.; Yang, Y. R. Enantioselective Total Synthesis of (-)-Alstoscholarisine a. J. Am. Chem. Soc. 2016, 138, 2560-2562. (65) Phukan, A.; Ganguli, J. N.; Dutta, D. K. ZnCl 2-Zn2+Montmorillonite Composite: Efficient Solid Acid Catalyst for Benzylation of Benzene. J. Mol. Catal. A-Chem. 2003, 202, 279-287.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TABLE OF CONTENTS

ACS Paragon Plus Environment

Page 10 of 10