Enhancement of catalytic activity by homo-dispersing S2O8 2 -Fe2O3 nanoparticles on SBA-15 through ultrasonic adsorption

Similar documents
Effect of lengthening alkyl spacer on hydroformylation performance of tethered phosphine modified Rh/SiO2 catalyst

The dynamic N1-methyladenosine methylome in eukaryotic messenger RNA 报告人 : 沈胤

Synthesis of anisole by vapor phase methylation of phenol with methanol over catalysts supported on activated alumina

Fabrication of ultrafine Pd nanoparticles on 3D ordered macroporous TiO2 for enhanced catalytic activity during diesel soot combustion

Ni based catalysts derived from a metal organic framework for selective oxidation of alkanes

Synthesis of PdS Au nanorods with asymmetric tips with improved H2 production efficiency in water splitting and increased photostability

A highly efficient flower-like cobalt catalyst for electroreduction of carbon dioxide

Mesoporous polyoxometalate based ionic hybrid as a highly effective heterogeneous catalyst for direct hydroxylation of benzene to phenol

Effect of promoters on the selective hydrogenolysis of glycerol over Pt/W containing catalysts

Integrating non-precious-metal cocatalyst Ni3N with g-c3n4 for enhanced photocatalytic H2 production in water under visible-light irradiation

A new approach to inducing Ti 3+ in anatase TiO2 for efficient photocatalytic hydrogen production

available at journal homepage:

Enhancement of the activity and durability in CO oxidation over silica supported Au nanoparticle catalyst via CeOx modification

Single-atom catalysis: Bridging the homo- and heterogeneous catalysis

Growth of Cu/SSZ 13 on SiC for selective catalytic reduction of NO

Species surface concentrations on a SAPO 34 catalyst exposed to a gas mixture

available at journal homepage:

Effect of the degree of dispersion of Pt over MgAl2O4 on the catalytic hydrogenation of benzaldehyde

Preparation of mesoporous Fe-Cu mixed metal oxide nanopowder as active and stable catalyst for low-temperature CO oxidation

Surface reactions of CuCl2 and HY zeolite during the preparation of CuY catalyst for the oxidative carbonylation of methanol

An efficient and stable Cu/SiO2 catalyst for the syntheses of ethylene glycol and methanol via chemoselective hydrogenation of ethylene carbonate

2. The lattice Boltzmann for porous flow and transport

In plasma catalytic degradation of toluene over different MnO2 polymorphs and study of reaction mechanism

Ionic covalent organic frameworks for highly effective catalysis

Selective oxidation of toluene using surface modified vanadium oxide nanobelts

Supplementary Information for

Hydrothermal synthesis of nanosized ZSM 22 and their use in the catalytic conversion of methanol

Ho modified Mn Ce/TiO2 for low temperature SCR of NOx with NH3: Evaluation and characterization

Resistance to SO2 poisoning of V2O5/TiO2 PILC catalyst for the selective catalytic reduction of NO by NH3

In situ preparation of mesoporous Fe/TiO2 catalyst using Pluronic F127 assisted sol gel process for mid temperature NH3 selective

Effect of Gd0.2Ce0.8O1.9 nanoparticles on the oxygen evolution reaction of La0.6Sr0.4Co0.2Fe0.8O3 δ anode in solid oxide electrolysis cell

SnO2 based solid solutions for CH4 deep oxidation: Quantifying the lattice capacity of SnO2 using an X ray diffraction extrapolation method

Pore structure effects on the kinetics of methanol oxidation over nanocast mesoporous perovskites

Catalytic performance and synthesis of a Pt/graphene TiO2 catalyst using an environmentally friendly microwave assisted solvothermal method

Highly selective hydrogenation of furfural to tetrahydrofurfuryl alcohol over MIL 101(Cr) NH2 supported Pd catalyst at low temperature

Design, Development and Application of Northeast Asia Resources and Environment Scientific Expedition Data Platform

Preparation of LaMnO3 for catalytic combustion of vinyl chloride

Magnetic Co/Al2O3 catalyst derived from hydrotalcite for hydrogenation of levulinic acid to γ-valerolactone

NiFe layered double hydroxide nanoparticles for efficiently enhancing performance of BiVO4 photoanode in

Ultrasonic synthesis of CoO/graphene nanohybrids as high performance anode materials for lithium ion batteries

Zinc doped g C3N4/BiVO4 as a Z scheme photocatalyst system for water splitting under visible light

Steering plasmonic hot electrons to realize enhanced full spectrum photocatalytic hydrogen evolution

Effects of Au nanoparticle size and metal support interaction on plasmon induced photocatalytic water oxidation

Synergetic effect between non thermal plasma and photocatalytic oxidation on the degradation of gas phase toluene: Role of ozone

Synthesis of Ag/AgCl/Fe S plasmonic catalyst for bisphenol A degradation in heterogeneous photo Fenton system under visible light irradiation

Enhanced visible light photocatalytic oxidation capability of carbon doped TiO2 via coupling with fly ash

Highly effective electrochemical water oxidation by copper oxide film generated in situ from Cu(II) tricine complex

Catalytic combustion of methane over Pd/SnO2 catalysts

La doped Pt/TiO2 as an efficient catalyst for room temperature oxidation of low concentration HCHO

Wheat flour derived N doped mesoporous carbon extrudes as an efficient support for Au catalyst in acetylene hydrochlorination

One step synthesis of graphitic carbon nitride nanosheets for efficient catalysis of phenol removal under visible light

N doped ordered mesoporous carbon as a multifunctional support of ultrafine Pt nanoparticles for hydrogenation of nitroarenes

SiO2 supported Au Ni bimetallic catalyst for the selective hydrogenation of acetylene

Biomolecule assisted, cost effective synthesis of a Zn0.9Cd0.1S solid solution for efficient photocatalytic hydrogen production under visible light

Visible light responsive carbon decorated p type semiconductor CaFe2O4 nanorod photocatalyst for efficient remediation of organic pollutants

Low cost and efficient visible light driven microspheres fabricated via an ion exchange route

Unsupported nanoporous palladium catalyzed chemoselective hydrogenation of quinolines: Heterolytic cleavage of H2 molecule

Electrocatalytic water oxidation by a nickel oxide film derived from a molecular precursor

沙强 / 讲师 随时欢迎对有机化学感兴趣的同学与我交流! 理学院化学系 从事专业 有机化学. 办公室 逸夫楼 6072 实验室 逸夫楼 6081 毕业院校 南京理工大学 电子邮箱 研 究 方 向 催化不对称合成 杂环骨架构建 卡宾化学 生物活性分子设计

available at journal homepage:

Silicoaluminophosphate molecular sieve DNL 6: Synthesis with a novel template, N,N' dimethylethylenediamine, and its catalytic application

Photocatalytic hydrogen evolution activity over MoS2/ZnIn2S4 microspheres

Promotional effects of Er incorporation in CeO2(ZrO2)/TiO2 for selective catalytic reduction of NO by NH3

Water oxidation catalytic ability of polypyridine complex containing a μ OH, μ O2 dicobalt(iii) core

d) There is a Web page that includes links to both Web page A and Web page B.

Carbon film encapsulated Fe2O3: An efficient catalyst for hydrogenation of nitroarenes

Novel structured Mo Cu Fe O composite for catalytic air oxidation of dye containing wastewater under ambient temperature and pressure

Rigorous back analysis of shear strength parameters of landslide slip

Photocatalytic Cr(VI) reduction and organic pollutant degradation in a stable 2D coordination polymer

Synthesis of novel p n heterojunction m Bi2O4/BiOCl nanocomposite with excellent photocatalytic activity through ion etching method

三类调度问题的复合派遣算法及其在医疗运营管理中的应用

Silver catalyzed three component reaction of phenyldiazoacetate with arylamine and imine

Synthesis of FER zeolite with piperidine as structure directing agent and its catalytic application

Facile preparation of composites for the visible light degradation of organic dyes

能源化学工程专业培养方案. Undergraduate Program for Specialty in Energy Chemical Engineering 专业负责人 : 何平分管院长 : 廖其龙院学术委员会主任 : 李玉香

Tuning the growth of Cu MOFs for efficient catalytic hydrolysis of carbonyl sulfide

Proton gradient transfer acid complexes and their catalytic performance for the synthesis of geranyl acetate

Effect of the support on cobalt carbide catalysts for sustainable production of olefins from syngas

Atomic & Molecular Clusters / 原子分子团簇 /

Photo induced self formation of dual cocatalysts on semiconductor surface

Cobalt nanoparticles encapsulated in nitrogen doped carbon for room temperature selective hydrogenation of nitroarenes

Microbiology. Zhao Liping 赵立平 Chen Feng. School of Life Science and Technology, Shanghai Jiao Tong University

High performance ORR electrocatalysts prepared via one step pyrolysis of riboflavin

Influence of surface strain on activity and selectivity of Pd based catalysts for the hydrogenation of acetylene: A DFT study

Effect of acidic promoters on the titania nanotubes supported V2O5 catalysts for the selective oxidation of methanol to dimethoxymethane

Synthesis and photocatalytic hydrogen production activity of the Ni CH3CH2NH2/H1.78Sr0.78Bi0.22Nb2O7 hybrid layered perovskite

A mini review on carbon based metal free electrocatalysts for

Preparation of a fullerene[60]-iron oxide complex for the photo-fenton degradation of organic contaminants under visible-light irradiation

Recent progress in Ag3PO4 based all solid state Z scheme photocatalytic systems

Catalytic effects of [Ag(H2O)(H3PW11O39)] 3 on a TiO2 anode for water oxidation

Highly enhanced visible-light photocatalytic hydrogen evolution on g-c3n4 decorated with vopc through - interaction

enzymatic cascade system

Promoting effect of polyaniline on Pd catalysts for the formic acid electrooxidation reaction

Chinese Journal of Applied Entomology 2014, 51(2): DOI: /j.issn 信息物质的化学分析技术 黄翠虹 , ; 2.

On the Quark model based on virtual spacetime and the origin of fractional charge

MASTER S DEGREE THESIS. Electrochemical Stability of Pt-Au alloy Nanoparticles and the Effect of Alloying Element (Au) on the Stability of Pt

Increasing the range of non noble metal single atom catalysts

Effects of composite oxide supports on catalytic performance of Ni-based catalysts for CO methanation

Coating Pd/Al2O3 catalysts with FeOx enhances both activity and selectivity in 1,3 butadiene hydrogenation

Service Bulletin-04 真空电容的外形尺寸

Transcription:

Chinese Journal of Catalysis 39 (2018) 955 963 催化学报 2018 年第 39 卷第 5 期 www.cjcatal.org available at www.sciencedirect.com journal homepage: www.elsevier.com/locate/chnjc Article Enhancement of catalytic activity by homo-dispersing S2O8 2 -Fe2O3 nanoparticles on SBA-15 through ultrasonic adsorption Qingyan Chu a,b, Jing Chen a, Wenhua Hou c, *, Haoxuan Yu d, Ping Wang a,d, Rui Liu a, Guangliang Song a, Hongjun Zhu a,#, Pingping Zhao d a College of Chemistry and Molecular Engineering, Nanjing Tech University, Nanjing 211816, Jiangsu, China b Research institute of QILU petrochemical company, SINOPEC, Zibo 255400, Shandong, China c Key Laboratory of Mesoscopic Chemistry of MOE, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023, Jiangsu, China d School of Chemical Engineering, Shandong University of Technology, Zibo 255049, Shandong, China A R T I C L E I N F O A B S T R A C T Article history: Received 23 November 2017 Accepted 29 December 2017 Published 5 May 2018 Keywords: Mesoporous superacid Nanoparticle Nano effect S2O8 2 -Fe2O3/SBA-15 Acidic site Ultrasonic adsorption Alcoholysis Mesoporous superacids S2O8 2 -Fe2O3/SBA-15 (SFS) with active nanoparticles are prepared by ultrasonic adsorption method. This method is adopted to ensure a homo-dispersed nanoparticle active phase, large specific surface area and many acidic sites. Compared with bulk S2O8 2 -Fe2O3, Brönsted acid catalysts and other reported catalysts, SFS with an Fe2O3 loading of 30% (SFS-30) exhibits an outstanding activity in the probe reaction of alcoholysis of styrene oxide by methanol with 100% yield. Moreover, SFS-30 also shows a more excellent catalytic performance than bulk S2O8 2 -Fe2O3 towards the alcoholysis of other ROHs (R = C2H5-C4H9). Lewis and Brönsted acid sites on the SFS-30 surfaces are confirmed by pyridine adsorbed infrared spectra. The highly efficient catalytic activity of SFS-30 may be attributed to the synergistic effect from the nano-effect of S2O8 2 -Fe2O3 nanoparticles and the mesostructure of SBA-15. Finally, SFS-30 shows a good catalytic reusability, providing an 84.1% yield after seven catalytic cycles. 2018, Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved. 1. Introduction Acid catalysts play an important role for various chemical reactions of industrial interest. Conventional acids (i.e., H2SO4, H3PO4 and p-toluenesulfonic acid) exhibit a high catalytic activity; however, significant risks are associated with their handling, containment, disposal as well as the issue of equipment corrosion [1,2]. For these reasons, new stringent environmental legislations against the use of these conventional acids in industrial applications have been implemented in many countries. Consequently, in recent years, solid superacids have received extensive attention because of their strong acidity, ease of separation, low corrosivity, stability, reusability and environmental friendliness [3 6]. Particularly, SO4 2 -MxOy superacids (i.e., SO4 2 -ZrO2, SO4 2 -TiO2 and SO4 2 -SnO2) have been the subject of many investigations because of their good catalytic performance and low waste generation [7,8]. Compared with SO4 2 -MxOy, S2O8 2 -MxOy with stronger acidities and catalytic activities have begun to attract interest. However, only a few kinds of MxOy compounds, such as ZrO2 and Al2O3, have been studied. Therefore, the exploration of metal oxides is of great value. In addition, Fe-doped superacids are well known for * Corresponding author. Tel/Fax: +86-25-58137480; E-mail: whou@nju.edu.cn # Corresponding author. E-mail: zhuhj@njtech.edu.cn This work was supported by the National Key Research and Development Program of China (2016YFB0301703), the Joint Innovation and Research Funding-prospective Joint Research Projects (BY2016005-06), and the National Natural Science Foundation of China (21506118). DOI: 10.1016/S1872-2067(17)63007-9 http://www.sciencedirect.com/science/journal/18722067 Chin. J. Catal., Vol. 39, No. 5, May 2018

956 Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 their good catalytic activity and stability in several industrial and environmental processes [9,10]. A research focus has been how to overcome the low surface area and porosity of bulk superacids to improve their specific surface area and enhance the catalytic activity of the active phase. Nowadays, with the development of nano-technology, various nanomaterials have been extensively used, including nanoparticles and nanocomposites, and have drawn considerable attention because of their high temperature sensitivity, high ductility, large surface area, high strain resistance, and low electrical resistivity [11,12]. How to turn materials into nanoparticles is a challenging topic. As an efficient and practical technology for the preparation of nano-materials, the ultrasonic technique has the advantages of introducing local stress intensities and ensuring a reproducible state of dispersions [13]. In addition, mesoporous silica materials have been identified as possible catalyst supports, particularly SBA-15, which have ordered hexagonally mesopores, thick channels, big pore diameters, and large surface areas [14 19]. In this context, we have developed homo-dispersed S2O8 2 -Fe2O3 (SF) nanoparticles supported on SBA-15 (S2O8 2 -Fe2O3/SBA-15, SFS) with a large specific surface area by ultrasonic adsorption method. The prepared mesoporous superacid showed better catalytic activities than bulk S2O8 2 -Fe2O3, Brönsted acid catalysts and the other catalysts reported for the probe reaction of alcoholysis of styrene oxide. In addition, the structure, surface acidity, and activity of the superacid catalysts were evaluated. 2. Experimental 2.1. Catalyst preparation All alcohols were analytical grade and purchased from SINOPHARM Chemical Reagent Co., Ltd. Styrene oxide (99%), (NH4)2S2O8 (98%) and Fe(NO3)3 9H2O (98%) were supplied from Energy Chemical. SBA-15 was obtained by Nanjing XFNANO Materials Tech Co., Ltd. The SFS solid superacids were prepared using the UA method (Fig. 1). The Fe(NO3)3 aqueous solution was slowly added to the SBA-15, then the mixture was evenly dispersed in an ultrasonic bath for 15 min. The Fe(NO3)3/SBA-15 was dried at 45 C for 12 h, and calcined at 550 C for 6 h to obtain the Fe2O3/SBA-15. During calcination, the temperature was increased at a rate of 1 C/min, starting from 50 C. A (NH4)2S2O8 aqueous solution was slowly added to the Fe2O3/SBA-15 to afford (NH4)2S2O8-Fe2O3/SBA-15, which was also dried at 45 C for 12 h and calcined at 500 C for 4 h after a heating ramp that started at 50 C with a rate of temperature increase of 2 C/min to obtain S2O8 2 -Fe2O3/SBA-15 (SFS-x, x = 10, 20, 25, 35, 40, 45). The Fe2O3 loading of x% was calculated according to Eq. (1). The Fe(NO3)3 9H2O/(NH4)2S2O8 molar ratio was 1:1. x(%) = WFe2O3/WSBA-15 100% (1) 2.2. Catalyst characterization X-ray fluorescence spectrometry (XRF) was recorded on a ZSX-100e. The surface morphology of the catalyst was observed using a FEI Sirion 200 field-emission scanning electron microscope (SEM) and a JEM-1200EX transmission electron microscope (TEM). The Fourier transform infrared (FT-IR) spectra were recorded on a Nicolet 360 FT-IR instrument (KBr discs) in the 4,000 400 cm 1 region. X-ray diffraction (XRD) patterns were collected on the X Pert Pro Multipurpose diffractometer using a Cu Kα radiation source at RT from 0.5 to 5.0 (small angle) and 5 to 50 (wide angle). Measurements were conducted using a voltage of 40 kv, and a current setting of 20 ma for small angle XRD and 40 ma for wide angle XRD. N2 adsorption-desorption isotherms and specific surface areas were measured at 196 C using a Micromeritics ASAP 2020 surface area and porosity analyzer. The X-ray photoelectron spectroscopy (XPS) measurements were performed on a RBD upgraded PHI-5000C ESCA system (Perkin-Elmer, Norwalk, CT, USA) with Mg Kα radiation (hν = 1253.6 ev), where binding energies were calibrated by using the containment carbon (C 1s = 284.6 ev). The infrared spectra of adsorbed pyridine (pyridine-ir spectra) were obtained on a PE Frontier FT-IR spectrometer [20]. 2.3. Catalytic tests A comparison of the catalytic performances of superacid catalysts tested in the probe reaction of alcoholysis of styrene oxide with MeOH was performed in a 50-mL flask (Scheme 1). Typically, styrene oxide (1.60 g, 13.3 mmol), MeOH (3.00 g, 93.6 mmol), and the catalyst (10.0 mg) were added to the reactor and stirred at 40 C. The reaction mixture was sampled periodically, and the samples were analyzed by gas chromatography (Agilent Technologies 7820A GC system, Palo Alto, CA, USA). Under the operating conditions adopted, the product was 2-methoxyl-2-phenylethanol with 100% selectivity, as identified by the analyses of 1 H NMR and 13 C NMR. The catalyst activity and stability were measured in detail. O OR Fig. 1. Well-ordered mesoporous S2O8 2 -Fe2O3/SBA-15 prepared by ultrasonic adsorption and dry-calcination. + ROH SF/SFS-30 (cat.) 40 o C Scheme 1. The probe reaction of alcoholysis of styrene oxide. OH

Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 957 3. Results and discussion 3.1. Catalyst characterization The morphology of the solid superacid SF was characterized by SEM and TEM. As shown from the SEM images (Fig. 2(a)), the crystal structure of SF was formed by the coordination of Fe2O3 and S2O8 2 at high temperature (Fig. 2(b)). Moreover, the lattice structures of SF could be clearly seen from the TEM images (Fig. 2(c) and 2(d)). The clear lattice fringes with two interval spacings of 0.405 and 0.30 nm corresponded to the crystallographic planes of monoclinic SF, in accordance with the XRD peaks at 29.7 o and 21.5 o (Fig. 4(c)). Elemental mappings were taken to gain insights into the distribution of SF (Fig. 2(e)). Furthermore, the XRF analysis showed the element content of SF was 41.8% O, 34.8% Fe, 22.1% S. The morphology analysis of SFS-30 is shown in Fig. 3. The external structure of SFS-30 (Fig. 3(a) and 3(b)) was similar to that of SBA 15. The TEM images clearly revealed the inner mesoporous channels in which SF nanoparticles of approximately 6 8 nm were uniformly distributed (Fig. 3(c)). Moreover, the clear lattice fringe with an interval spacing of 0.44 nm (Fig. 3(d)) corresponded to the crystallographic planes of monoclinic SFS-30, in agreement with the XRD peak at 20.3 o (Fig. 4(d)). To gain insights into the distribution of SFS-30, elemental mappings were taken (Fig. 3(e)). It could be seen that the four elements of O, Fe, S and Si were evenly distributed in the prepared catalytic material, which showed that the active components were uniformly dispersed in SBA-15, and no aggregation occurred. Furthermore, the XRF analysis showed the element content of SFS-30 was 42.9% O, 18.5% Fe, 5% S, and 32.7% Si. The FT-IR spectra of Fe2O3 and SF are shown in Fig. 4(a). Compared with Fe2O3, SF showed the emergence of new bands Fig. 2. (a) SEM images of SF; (b d) TEM images of SF; TEM images (e) and elemental mapping (f) of the composite SF. Fig. 3. (a) SEM images of SFS-30; (b d) TEM images of SFS-30; TEM images (e) and elemental mapping (f) of the composite SFS-30. in the range of 900 1300 cm 1, which were attributed to the S =O or S O bonds in the acid structures of the catalysts [21,22]. In this range, the peaks at 995, 1113 and 1127 cm 1 were assigned to the stretching vibration of S O. Meanwhile, the band at 1227 cm 1 was assigned to the stretching vibration of S=O. The attribution of these characteristic vibration peaks was in accordance with the reported S2O8 2 ion of solid superacids [2,23 26]. This result indicated a very strong interaction of surface persulfate species with Fe(III) species, which was the significant reason for the formation of the active acid center on the surface of SF [27,28]. The FT-IR spectra characterizing SBA-15 and SFS-30 are shown in Fig. 4(b). The band at 964 cm 1 was related to the characteristic stretching vibration of Si-OH of SBA-15 [29,30]. It was apparent that the absorption band at 964 cm 1 moved to 968 cm 1 and became weaker after loading SF on the pore surface of SBA-15. This effect was attributed to the stretching vibration of Si OH being perturbed by Fe 3+ in a neighboring position [28,31]. The special bands in the range of 900 1300 cm 1, attributed to the S=O or S O bonds, were covered by the band at 1091 cm 1, which was attributed to the asymmetric stretching vibration of Si O Si in the SBA-15 [32,33]. However, it was obvious that the band at 1097 cm 1 broadened compared with the band at 1091 cm 1, which indicated the successful loading of SF nanoparticles on the SBA-15. The XRD patterns of Fe2O3 and SF in the wide-angle region of 3 60 are shown in Fig.4(c). The prepared Fe2O3 exhibited all the well-resolved characteristic peaks of the Fe2O3 crystallite (JCPDS card No. 33-0664) with the six peaks displayed at 2θ = 24.1, 33.1, 35.6, 40.8, 49.4 and 54.0. Six new peaks ap-

958 Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 100 (a) 100 964 (b) Transmittance (%) 80 60 40 20 0 1270 1238 995 1073 1085 Transmittance (%) 80 60 40 20 0 SBA-15 SFS-30 1097 968 1091 Fe 2 O 3 1127 1113 SF 3500 3000 2500 2000 1500 1000 500 3500 3000 2500 2000 1500 1000 500 Wavenumbers (cm 1 ) Wavenumber (cm 1 ) Fe 2 O 3 SF (c) (d) Intensity (a.u.) Intensity (a.u.) Intensity (a.u.) 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 2 /( O ) SFS-30 SBA-15 10 20 30 40 50 60 2 /( ) O 10 20 30 40 50 60 2 /( ) Fig. 4. FT-IR spectra of Fe2O3, SF (a), SBA-15 and SFS-30 (b); XRD of Fe2O3, SF (c), SBA-15 and SFS-30 (d). peared in the diffraction pattern of SF after (NH4)2S2O8 adsorption and calcination. The six peaks displayed at 2θ = 14.7, 20.2, 21.5, 24.7, 29.7 and 32.5 were well matched with the standard JCPDS card No. 33-0679, and resulted from the interaction of Fe2O3 with (NH4)2S2O8. The XRD patterns of SBA-15 and SFS-30 in the small and wide-angle regions are shown in Fig. 4(d). The small angle (0.4 10 ) XRD patterns were similar for SBA-15 and SFS-30 and showed that S2O8 2 -Fe2O3 was very well dispersed and did not damage the mesoporous structure of SBA-15. However, there were significant differences in the wide-angle XRD pattern (10 60 ). SFS-30 clearly exhibited six new peaks at 2θ = 14.7, 20.3, 21.5, 24.7, 29.7 and 32.5 (JCPDS card No. 33-0679), which corresponded to the XRD of SF, and which did not appear in the XRD of SBA-15. Such results indicated the successful loading of SF nanoparticles on SBA-15. To investigate the chemical states of Fe and S in the samples, XPS measurements were performed. Fig. 5(a) and 5(b) shows the Fe 2p and S 2p XPS spectra of SFS-30. The SFS-30 contained two peaks at 706.75 and 720.07 ev (Fig. 5(a)), which could be assigned to the Fe 3+ species. The S 6+ species was revealed by the binding energy of 169 ev in Fig. 5(b) [26]. The N2 adsorption-desorption isotherms of SFS-30 are shown in Fig. 5(c). According to IUPAC classifications [34], the isotherms exhibited a type IV pattern and both showed a clear loop (H1 hysteresis loop) in the relative pressure range of 0.4 0.8. The presence of a hysteresis loop, which arose from capillary condensation, indicated the presence of regular mesoporous channels in the materials. This further confirmed that SFS-30 retained the intact and ordered mesoporous structure of SBA-15 after S2O8 2 -Fe2O3 loading. The parameters describing the textural properties and catalytic activities of bulk SF and SFS-30 are summarized in Table 1. The BET surface area and pore volume of SFS-30 were obviously larger than those of SF. In addition, Fig. 5(c) shows the N2 adsorption-desorption isotherms of SFS-40, where the BET specific surface area, pore volume, and pore size were 218.09 m 2 /g, 0.21 cm 3 /g and 3.50 nm, respectively. The pyridine-ir spectra were employed to evaluate the type and strength of Brönsted and Lewis acid sites on SFS-30 [35]. The pyridine-ir spectra of SFS-30 (Fig. 5(d)) exhibited bands at 1449 and 1609 cm 1, which demonstrated the existence of Lewis acid sites in the materials. The bands at 1543 and 1639 cm 1 were those characteristic of the pyridinium ion, which showed the presence of Brönsted acid sites. The band at 1490 cm 1 was a combination between two separate bands, namely, those at 1449 and 1543 cm 1, which corresponded to Brönsted and Lewis acid sites, respectively. These indicated that both Brönsted and Lewis acid sites existed in the sample. In addition, the changes of Brönsted acidity and Lewis acidity were observed from the Fig. 5(d), for desorption temperatures of 150, 250, 350, and 450 C. Compared with Lewis acidity, the

Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 959 Counts (s) 26000 24000 22000 20000 18000 16000 14000 12000 10000 (a) Fe 2p 2p 3/2 706.75 13.32 2p 1/2 SF SFS-30 8000 700 710 720 730 740 Volume adsoreb ( cm 3 /g, STP ) (c) Binding energy (ev) 0.0 0.2 0.4 0.6 0.8 1.0 Relative pressure ( p/p 0 ) SFS-30 SFS-40 Counts (s) 18000 16000 (b) SF S 2p 169 SFS-30 14000 12000 10000 8000 6000 4000 2000 0 158 160 162 164 166 168 170 172 174 176 1449 (LA) Binding energy (ev) 1490 (LA+BA) 1543 (BA) 1609 (LA) 1639 (BA) 1400 1450 1500 1550 1600 1650 1700 1750 Wavenumber (cm 1 ) 150 o C 250 o C 350 o C 450 o C Fig. 5. Fe 2p (a) and S 2p (b) XPS spectra of SF and SFS-30; (c) N2 adsorption-desorption isotherms of SFS-30 and SFS-40; (d) The pyridine-ir spectra of SFS-30. Absorbance (d) Brönsted acidity was significantly reduced with increasing temperature. 3.2. Influence of loading amounts of Fe2O3 on SBA-15 To understand the influence of loading Fe2O3 on the mesoporous solid superacid, the catalytic performance of SFS-x (x = 10, 20, 25, 30, 35, 40, 45) in the alcoholysis was investigated. As shown in Fig. 6(a), the activity of the catalyst increased as the enhancement of Fe2O3 loading amount passed through a maximum at x = 30 and then declined. This implied that when x was below 30, the surface active acid sites increased with the loading of SF. However, when x was above 30, the excessive SF probably caused a partial aggregation of catalyst, and hindered the pore diffusion and decreased the catalytic activity. It indicated that the increasing loading led to a sustained decrease of Table 1 BET surface area, textural data and catalytic activities for bulk SF, SFS-30 and SBA-15. Sample ABET a (m 2 /g) Vp b (cm 3 /g) D c (nm) Catalytic activity d yield (%) reaction time (min) Bulk SF 6.57 0.02 11.33 99.5, 45 SFS-30 263.78 0.38 6.95 100, 30 SBA-15 605.47 0.86 7.01 a ABET = BET surface area. b Vp = Pore volume. c D = Pore size. d Reaction conditions were same to Fig. 6(b). the active surface area and mesopore volume, which had a significant influence on the properties of the acid, as has been previously reported [36]. 3.3. Comparison of catalytic activity A comparison between the catalytic activity of SF and SFS-30 was performed and the results are shown in Fig. 6(b) and (c). When using 10.0 mg of SF catalyst, it was found that the optimal molar ratio of styrene oxide to MeOH (93.6 mmol) was 1:7. As shown in Fig. 6(b), the reaction yield of 99.5% at 40 C required 45 min for SF. When using the SFS-30 catalyst, the reaction easily achieved a yield of 100% after 30 min. Next, we investigated the alcoholysis of styrene oxide with various other alcohols with SFS-30. As illustrated in Fig. 6(c), it is worth mentioning that the alcoholysis yields of the SFS-30 catalyzed reactions were 99.7% and 99.6% with EtOH and n-proh, respectively. However, the yields of the same reactions when catalyzed by SF were only 95.4% and 69.5%. Furthermore, the catalytic activity of SFS-30 was higher than that of SF in the alcoholysis of styrene oxide with n-buoh, i-proh, i-buoh and s-buoh. From these results, we can conclude that the catalytic activity of SFS-30 was higher than that of bulk SF, which indicated that the SF nanoparticles loading on the SBA-15 not only enhanced the active surface area of the activity components (Table 1), but also reduced the amounts of effective compo-

960 Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 Fig. 6. (a) Influence of load amounts of Fe2O3 on SBA-15 for alcoholysis of styrene oxide with MeOH, 15 min; (b) Comparison between the catalytic activity of SF and SFS-30, catalyst (10.0 mg), styrene oxide (1.60 g, 13.3 mmol), MeOH (3.00 g, 93.6 mmol), 40 C; (c) SF and SFS-30 catalyzed alcoholysis of styrene oxide with other alcohols, catalyst (10.0 mg), styrene oxide (1.60 g, 13.3 mmol), alcohols (93.6 mmol), 40 C, 90 min; (d) Catalytic reusability of SFS-30, catalyst (50.0 mg), styrene oxide (8.00 g, 66.5 mmol), MeOH (15.00 g, 468 mmol), 40 C, 30 min. nents. With the synergistic effect from the nano effect of SF nanoparticles and the mesostructure of SBA-15, the reaction time of catalyzed SFS-30 was one-third lower than that required when using bulk SF. The catalytic performance of SFS-30 and several different Brönsted acids for the alcoholysis is compared in Table 2. Similar to the non-catalyzed system, Fe2O3/SBA-15 did not yield any products. Obviously, the styrene-based sulfonic resins presented a lower yield with 12.2% conversion and 89% selectivity. H2SO4 and p-toluenesulfonic acid exhibited good catalytic properties close to those of SFS-30 with selectivities of 98.9% and 98.8%, respectively. However, the two acids were dissolved in the reaction mixtures and were difficult to isolate and reuse: alkaline materials must be used to neutralize the acids to avoid a continuous increase of by-products (1,2-dimethoxyl-1-phenylethane). As a result, large amounts of salt and wastewater will be produced, which represents an environmental issue. By contrast, SFS-30 exhibited an excellent activity with 100% yield and an ease of separation. However, Fe2O3/SBA-15 did not yield any products, which was attributed to more Lewis acidity sites of SFS-30, which originated from the cooperation of S2O8 2 and Fe2O3 [37]. A comparison between the catalytic activity of our SFS-30 and that of other reported catalysts is listed in Table 3. SFS-30 presented the highest catalytic activity, even with the highest substrate/catalyst (S/C) ratio of 160:1 and the lowest MeOH/styrene oxide molar ratio of 7:1. It was also clear that S2O8 2 -Fe2O3 showed a good catalytic performance in the reaction (Fig. 6(b)). [V IV (TPP)(OTf)2] was the second highest catalytic activity, but the high activity could only be obtained with an S/C of 24:1 and a very high MeOH/styrene oxide molar ratio of 98.8:1. As a result, SFS-30 with the synergistic effect between the nano effect of S2O8 2 -Fe2O3 nanoparticles and the mesostructure of SBA-15 exhibited the more highly efficient Table 2 Catalytic performances of SFS-30 and several different Brönsted acids for the alcoholysis of styrene oxide with MeOH. Catalyst Conversion (%) Selectivity (%) 0 0 Fe2O3/SBA-15 0 0 S2O8 2 -Fe2O3/SBA-15 * 100 100 H2SO4 99.5 98.9 p-toluenesulfonic acid 99 98.8 Styrene-based sulfonic resins 12.2 89 Reaction conditions: catalyst (10.0 mg), styrene oxide (1.60 g, 13.3 mmol), MeOH (3.00 g, 93.6 mmol), 40 C, 45 min. * Reaction time was 30 min.

Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 961 Table 3 Comparison between the catalytic activity of SFS-30 and that of other reported catalysts for the alcoholysis of styrene oxide with MeOH. Catalyst catalytic activity compared with other reported catalysts. 3.4. Catalyst reusability Since the reusability is an important factor influencing the practical applications of catalysts, we carried out a seven run test over the SFS-30 solid superacid under the optimized reaction conditions to evaluate the recyclability of the catalyst (Fig. 6(d)). The selectivity of alcoholysis remained at 100% in the seven runs test over the SFS-30. Compared with the fresh catalyst, a slow decline in the conversion was observed after seven runs for the SFS-30 catalyzed alcoholysis reaction; however, the yield still remained high at 84.1%. The reason for the decrease of the conversion after reuse was mainly the chemical deactivation of SFS-30 arising from active phase shedding. This result indicated that the SFS-30 exhibited a good stability and can potentially be used in industrial applications. 4. Conclusions MeOH/Strene oxide molar ratio Substrate/ Catalyst (S/C) Yield (%) Ref. S2O8 2 -Fe2O3/SBA-15 7:1 160:1 100 This work [V IV (TPP)(OTf)2] 98.8:1 24:1 100 [37] MS-SO3H 37:1 6:1 100 [38] Fe(Cp)2BF4 61.8:1 20:1 99 [39] Fe(Cp)2PF6 61.8:1 20:1 99 [39] ZrCl4 61.8:1 20:1 98 [39] (NH4)2Ce(NO3)6 61.8:1 20:1 91 [39] FeCl3 61.8:1 20:1 99 [39] InCl3 61.8:1 20:1 81 [39] [Fe(BTC)] 61.8:1 4.8:1 48 [40] [Cu3(BTC)2] 61.8:1 4.8:1 24 [40] [Al2(BDC)3] 61.8:1 4.8:1 20 [40] Iron citrate 61.8:1 4.8:1 24 [40] Fe2(NO3)2 9H2O 61.8:1 2.4:1 97 [40] Sulfated Zr-doped [TiNbO5]- nanoplates 49.4:1 2.4:1 99 [41] ZrO(NO3)2 nh2o 74:1 5:1 98 [42] Well-dispersed S2O8 2 -Fe2O3/SBA-15 mesostructures with active nanoparticles, prepared by ultrasonic adsorption, present a larger specific surface area and larger mesopore volume with surface acid sites than bulk S2O8 2 -Fe2O3. According to our experimental results, S2O8 2 -Fe2O3/SBA-15 with a 30% Fe2O3 loading exhibited the highest activity in the alcoholysis of styrene oxide with MeOH compared with bulk S2O8 2 -Fe2O3 and other reported Brönsted acids, Lewis acids and other catalysts. Furthermore, this mesoporous superacid also exhibited a more excellent highly efficient catalysis for alcoholysis with other ROHs (R = C2H5-C4H9) than bulk S2O8 2 -Fe2O3 and better reusability of the catalyst. Therefore, mesoporous S2O8 2 -Fe2O3/SBA-15 possessing highly efficient catalytic activities, good stability and economy shows great promise for industrial application. References [1] X. Z. Liang, Chem. Eng. J., 2015, 264, 251 257. [2] Q. Zhao, S. M. Meng, J. L. Wang, Z. P. Li, J. H. Liu, Y. Guo, Ceram. Int., 2014, 40, 16183 16187. [3] L. Qin, G. L. Zhang, Z. Fan, Y. J. Wu, X. W. Guo, M. Liu, Chem. Eng. J., 2014, 244, 296 306. [4] G. D. Yadav, S. B. Kamble, Appl. Catal. A, 2012, 433, 265 274. [5] Z. K. Zhao, J. F. Ran, Appl. Catal. A, 2015, 503, 77 83. [6] W. Li, K. B. Chi, H. Liu, H. J. Ma, W. Qu, C. X. Wang, G. Lv, Z. J. Tian, Appl. Catal. A, 2017, 537, 59 65. [7] Y. Fang, Y. Li, Q. Zhang, X. F. Sun, H. X. Fan, N. Xu, L. Gang, Carbohyd. Polym., 2015, 131, 9 14. [8] L. Gu, X. Chen, Y. Zhou, Q. L. Zhu, H. F. Huang, H. F. Lu, Chin. J. Catal., 2017, 38, 607 615. [9] F. H. Alhassan, U. Rashid, Y. H. Taufiq-Yap, Fuel, 2015, 142, 38 45. [10] H. J. Li, H. L. Song, L. W. Chen, C.G. Xia, Appl. Catal. B, 2015, 165, 466 476. [11] M. Dastkhoon, M. Ghaedi, A. Asfaram, A. Goudarzi, S. M. Mohammadi, S. B. Wang, Ultrason Sonochem., 2017, 37, 94 105. [12] R. Y. Li, F. P. Xiao, S. Amirkhanian, Z. P. You, J. Huang, Constr. Build. Mater., 2017, 143, 633 648. [13] R. R. Retamal Marín, F. Babick, M. Stintz, Powder Technol., 2017, 318, 451 458. [14] F. Z. El Berrichi, C. Pham-Huu, L. Cherif, B. Louis, M. J. Ledoux, Catal. Commun., 2011, 12, 790 793. [15] M. Zare, Z. Moradi-Shoeili, F. Ashouri, M. Bagherzadeh, Catal. Commun., 2017, 88, 9 12. [16] R. J. Kalbasi, M. Kolahdoozan, K. Shahabian, F. Zamani, Catal. Commun., 2010, 11, 1109 1115. [17] Y. Qin, Z. P. Qu, C. Dong, N. Huang, Chin. J. Catal., 2017, 38, 1603 1612. [18] H. J. Shen, X. Y. Wu, D. H. Jiang, X. N. Li, J. Ni, Chin. J. Catal., 2017, 38, 1597 1602. [19] K. N. Tayade, L. Y. Wang, S. S. Shang, W. Dai, M. Mishra, S. Gao, Chin. J. Catal., 2017, 38, 758 766. [20] Z. C. Miao, H. L. Song, H. H. Zhao, L. L. Xu, L. J. Chou, Catal. Sci. Technol., 2014, 4, 838 850. [21] O. V. Manoilova, R. Olindo, C. O. Areán, J. A. Lercher, Catal. Commun., 2007, 8, 865 870. [22] P. F. Chen, M. X. Du, H. Lei, Y. Wang, G. L. Zhang, F. B. Zhang, X. B. Fan, Catal. Commun., 2012, 18, 47 50. [23] J. X. Wang, H. Pan, A. Q. Wang, X. Y. Tian, X. L. Wu, Y. Q. Wang, Catal. Commun., 2015, 62, 29 33. [24] H. Song, L. L. Zhao, N. Wang, Chin. J. Chem. Eng., 2016, 25, 74 78. [25] G. P. Fan, M. Shen, Z. Zhang, F. R. Jia, J. Rare Earth, 2009, 27, 437 442. [26] H. Song, L. L. Zhao, N. Wang, F. Li, Appl. Catal. A, 2016, 526, 37 44. [27] Y. J. Wu, L. Qin, G. L. Zhang, L. Chen, X. W. Guo, M. Liu, Ind. Eng. Chem. Res., 2013, 52, 16698 16708. [28] B. B. Chang, J. Fu, Y. L. Tian, X. P. Dong, Appl. Catal. A, 2012, 437, 149 154. [29] W. J. Ding, W. S. Zhu, J. Xiong, Y. Lei, A. M. Wei, Z. Ming, H. M. Li, Chem. Eng. J., 2015, 266, 213 221. [30] P. M. Rao, A. Wolfson, S. Kababya, S. Vega, M. V. Landau, J. Catal., 2005, 232, 210 225. [31] S. C. Laha, P. Mukherjee, S. R. Sainkar, R. Kumar, J. Catal., 2002, 207, 213 223. [32] S. G. Wang, K. K. Wang, C. Dai, H. Z. Shi, J. L. Li, Chem. Eng. J., 2015, 262, 897 903.

962 Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 Graphical Abstract Chin. J. Catal., 2018, 39: 955 963 doi: 10.1016/S1872-2067(17)63007-9 Enhancement of catalytic activity by homo-dispersing S2O8 2 -Fe2O3 nanoparticles on SBA-15 through ultrasonic adsorption Qingyan Chu, Jing Chen, Wenhua Hou *, Haoxuan Yu, Ping Wang, Rui Liu, Guangliang Song, Hongjun Zhu *, Pingping Zhao Nanjing Tech University; Research institute of QILU petrochemical company, SINOPEC; Nanjing University; Shandong University of Technology Mesoporous superacids S2O8 2 -Fe2O3/SBA-15 with active nanoparticles are prepared by ultrasonic adsorption. TEM images show S2O8 2 -Fe2O3 nanoparticles of approximately 6 8 nm are uniformly distributed in the inner mesoporous channels of SBA-15. [33] P. Shukla, W. B. Shao, H. Q. Sun, H. M. Ang, M. Tadé, Chem. Eng. J., 2010, 164, 255 260. [34] K. S. W. Sing, D. H. Ererett, R. A. W. Haul, L. Moscou, R. A. Pierotti, J. Rouquerol, T. Siemieniewska, Pure Appl. Chem., 1985, 57, 603 619. [35] V. Krishna, G. Naresh, V. V. Kumar, R. Sarkari, A. H. Padmasri, A. Venugopal, Appl. Catal. B, 2016, 193, 58 66. [36] S. Garg, K. Soni, G. M. Kumaran, R. Bal, K. Gora-Marek, J. K. Gupta, L. D. Sharma, G. Murali Dhar, Catal. Today, 2009, 141, 125 129. [37] S. A. Taghavi, M. Moghadam, I. Mohammadpoor-Baltork, S. Tangestaninejad, V. Mirkhani, A. R. Khosropour, V. Ahmadi, Polyhedron, 2011, 30, 2244 2252. [38] Y. H. Kang, X. D. Liu, N. Yan, Y. Jiang, X. Q. Liu, L. B. Sun, J. R. Li, J. Am. Chem. Soc., 2016, 138, 6099 6102. [39] G. D. Yadav, S. Singh, Tetrahedron Lett., 2014, 55, 3979 3983. [40] A. Dhakshinamoorthy, M. Alvaro, H. Garcia, Chem. Eur. J., 2010, 16, 8530 8536. [41] L. H. Zhang, C. H. Hu, W. G. Mei, J. F. Zhang, L. Y. Cheng, N. H. Xue, W. P. Ding, J. Chen, W. H. Hou, Appl. Surf. Sci., 2015, 357, 1951 1957. [42] S. S. Shinde, M. S. Said, T. B. Surwase, P. Kumar, Tetrahedron Lett., 2015, 56, 5916 5919. 超声吸附制备均匀分散在 SBA-15 的 S 2 O 8 2 -Fe 2 O 3 纳米粒子 : 增强 S 2 O 8 2 -Fe 2 O 3 本体催化活性 楚庆岩 a,b, 陈静 a, 侯文华 c,*, 余昊轩 d, 王平 a,d, 刘睿 a, 宋广亮 a a,# d, 朱红军, 赵萍萍 a 南京工业大学化学与分子工程学院, 江苏南京 211816 b 中国石化齐鲁分公司研究院, 山东淄博 255400 c 南京大学化学化工学院介观化学教育部重点实验室, 江苏南京 210023 d 山东理工大学化工学院, 山东淄博 255049 摘要 : 酸催化剂在化学反应和化工生产中具有重要的作用. 传统无机酸, 如 H 2 SO 4, H 3 PO 4 和对甲苯磺酸等具有较高的催化活性, 但是存在污染大 设备腐蚀严重以及催化剂不能重复使用等问题. 固体酸具有酸性强 易分离 环境友好以及稳定性和重复使用性好等特点因而近年来越来越引起人们的关注. 其中, SO 2 4 -M x O y 固体超强酸 ( 如 SO 2 4 -ZrO 2, SO 2 4 -TiO 2 和 SO 2 4 -SnO 2 等 ) 因具有很好的催化性能而备受关注. 相比 SO 2 4 -M x O y, S 2 O 2 8 -M x O y 具有更强的酸性和稳定性而成为研究的重点. 如何克服固体超强酸本体的低比表面积和孔容, 增加其比表面积和催化活性是固体超强酸研究的热点. 超声吸附法可保证所制介孔固体酸活性组分均匀分散, 以及大的比表面积和更多的酸性位点. 因此采用超声吸附法制备了一种新型介孔固体酸 S 2 O 2 8 -Fe 2 O 3 /SBA-15. 相比 S 2 O 2 8 -Fe 2 O 3 本体 B 酸和文献报道催化剂, 负载 30%Fe 2 O 3 的 S 2 O 2 8 -Fe 2 O 3 /SBA-15 在环氧苯乙烷甲醇醇解的探针反应中显示出很高的催化活性, 反应收率为 100%. S 2 O 2 8 -Fe 2 O 3 纳米粒子的纳米效应和 SBA-15 介孔结构的协同作用使 S 2 O 2 8 -Fe 2 O 3 /SBA-15 具有高催化活性. 相比 S 2 O 2 8 -Fe 2 O 3 本体, 采用超声分散技术制备的 S 2 O 2 8 -Fe 2 O 3 /SBA-15 固体超强酸具有典型的介孔结构 大的比表面积

Qingyan Chu et al. / Chinese Journal of Catalysis 39 (2018) 955 963 963 和孔容, 并且表面富含酸性位点. 并且吡啶红外分析 S 2 O 2 8 -Fe 2 O 3 /SBA-15 表面富含 L 酸和 B 酸. 环氧苯乙烷甲醇醇解探针反应表明, Fe 2 O 3 负载量为 30% 时, S 2 O 2 8 -Fe 2 O 3 /SBA-15 的催化活性最高, 优于 S 2 O 2 8 -Fe 2 O 3 本体和已报道的布朗酸和路易斯酸等催化剂, 将醇底物拓展 (ROHs, R = C 2 H 5 -C 4 H 9 ), S 2 O 2 8 -Fe 2 O 3 /SBA-15 的催化活性也优于 S 2 O 2 8 -Fe 2 O 3 本体. 同时, S 2 O 2 8 -Fe 2 O 3 /SBA-15 具有很好的重复使用性能, 连续使用七次, 反应收率在 84.1% 以上. 总之, 具有高催化活性 好的稳定性和经济性的 S 2 O 2 8 -Fe 2 O 3 /SBA-15 具有广阔的应用前景. 关键词 : 介孔固体酸 ; 纳米粒子 ; 纳米效应 ; S 2 O 2 8 -Fe 2 O 3 /SBA-15; 酸性位 ; 超声分散 ; 醇解 收稿日期 : 2017-11-23. 接受日期 : 2017-12-29. 出版日期 : 2018-05-05. * 通讯联系人. 电话 / 传真 : (025)58137480; 电子信箱 : whou@nju.edu.cn # 通讯联系人. 电子信箱 : zhuhj@njtech.edu.cn 基金来源 : 国家重点研究发展计划项目 (2016YFB0301703); 江苏省产学研前瞻性联合研究项目 (BY2016005-06); 国家自然科学基金 (21506118). 本文的电子版全文由 Elsevier 出版社在 ScienceDirect 上出版 (http://www.sciencedirect.com/science/journal/18722067).