Multifunctional Tin-Based Heterogeneous Catalyst for Catalytic

Feb 21, 2018 - The column oven was set to 65 °C. The column can efficiently resolute glucose and fructose. HMF was detected using a Diamonsil C18 col...
4 downloads 17 Views 3MB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

Multifunctional Tin-based heterogeneous catalyst for catalytic conversion of glucose to 5-hydroxymethylfurfural Ke Li, Mengmeng Du, and Peijun Ji ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b00745 • Publication Date (Web): 21 Feb 2018 Downloaded from http://pubs.acs.org on February 26, 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.

ACS Sustainable Chemistry & Engineering 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 32 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 Sustainable Chemistry & Engineering

Multifunctional Tin-based heterogeneous catalyst for catalytic conversion of glucose to 5-hydroxymethylfurfural Ke Li†, Mengmeng Du†, Peijun Ji* Department of Chemical Engineering, Beijing University of Chemical Technology, Beisanhuandonglu 15, Chaoyang district, Beijing, 100029, China E-mail: [email protected] (P. Ji) ABSTRACT Using 5-sulfoisophthalic acid as the ligand, Tin porous coordination polymer (SnPCP) was synthesized on polydopamine-coated MnO2 (MnO2−PDA). The novel composite SnPCP@MnO2−PDA

was

used

for

conversion

of

glucose

into

5-hydroxymethylfurfural (HMF). The tetrahedral coordinated tin and the sulfonic groups of the ligand catalyze glucose isomerization to fructose and fructose dehydration to HMF, respectively. Thus, the composite is a bifunctional catalyst. The porous structure of SnPCP of the composite facilitates the transport of glucose, intermediate and HMF within the catalyst. In addition, MnO2−PDA was found to be able to catalyze the conversion of glucose to HMF. The synergistic effect of SnPCP and MnO2−PDA achieved HMF yields of 55.8% in DMSO and 41.2% in water/THF. Consecutive use of SnPCP@MnO2−PDA demonstrated that after 5 cycles, the activity loss is not significant in terms of the HMF yield and glucose conversion. KEYWORDS: Tin-based catalyst; 5-hydroxymethylfurfural; glucose; synergistic effect.

 INTRODUCTION 5-hydroxymethylfurfural (HMF) is a

biomass-based important platform

compound which can be used to synthesize biofuel and high-value chemicals applicable in pharmaceuticals, petroleum industry and furanose-based polymers.1-4

*

Corresponding author. E-mail: [email protected] 1

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 2 of 32

Glucose is the most attractive biomass-derived carbohydrate. It can be used as starting material for producing HMF.5 Producing the platform chemical HMF from glucose contributes to the sustainable development. When conventional Brønsted acids were used to catalyze the dehydration of glucose into HMF, HMF was obtained with a low yield, because the stability of the glucose ring makes processing difficult.6,7 Therefore effective catalysts are required for the synthesis of HMF from glucose. Producing HMF from glucose can be carried out in two steps, glucose is isomerized to fructose and fructose is dehydrated to HMF, as illustrated in Scheme 1.7

Scheme 1. Simplified reaction scheme for conversion of glucose to HMF through glucose isomerization and subsequent dehydration.

Due to the economic prospect producing HMF from glucose, many investigations have been performed. Chemical catalysts, homogeneous and heterogeneous, have been investigated for catalyzing glucose isomerization, such as hydrotalcite catalyst,8 cationized zeolites,9,10 organic bases,11 anionized resins,12 and phosphoric acid treated metal oxides.13 Tin based catalysts, including Lewis acidic metal salt SnCl4 and Sn-Beta,14-24 are promising catalysts for carbohydrate isomerization. Liquid acids, such as H2SO4, HCl in water, and H3PO4, can be used to catalyze fructose dehydration. There are some concerns for using these homogeneous catalysts, such as environmental pollution and difficult in product separation and recovery of catalysts .25 In contrast, solid acid catalysts are easy to recover and cause less environmental pollution.26,27 Sulfonate groups functionalized solid acids are well known to be favorable for fructose dehydration.25,28,29 2

ACS Paragon Plus Environment

Page 3 of 32 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 Sustainable Chemistry & Engineering

Isomerization of glucose into fructose catalyzed by glucose isomerase is an equilibrium-limited reaction. Simeonov and coworkers used two-pot approach to convert glucose to HMF by using biocatalyst and chemical catalyst separately, isomerization of glucose to fructose by glucose isomerase and fructose dehydration to HMF by Brønsted acids.30 However, the operation conditions are limited when using glucose isomerase, as the enzyme is denatured at high temperature and harsh pH conditions.31-32 On the contrary, Lewis acids solid catalysts, which have shown to be active for glucose isomerization, can catalyze at a wide range of operation conditions.31-32 The catalysts possessing both Lewis/Brønsted acids sites have been investigated for one-pot conversion of glucose to HMF.31-33 The one-pot approach does not require separating and purifying the intermediates, and can get around the reaction equilibrium of glucose isomerization. A polydivinylbenzene polymeric material was used for catalyzing glucose conversion to HMF in a two-phases system, HMF with a stable yield was produced.32 Sulfonated MIL-101Cr was applied for producing HMF from glucose, in the mixed solvents THF/H2O, a glucose conversion of 29% after 24 h reaction was achieved.33 In another one-pot approach, Lewis acidic Sn-Beta and HCl were used as the catalyst, and the HMF yields were in excess of 50%.7 These results showed that combining Brønsted acid and Lewis acid can enhance glucose conversion to HMF. SnCl4 dissolved in EMIMBr could efficiently convert high-concentration glucose into HMF. 15 Sn-Beta has shown promising catalytic properties on lab-scale, however, the use of fluoride ions and relatively long crystallization times for synthesizing this material might hinder its industrial application.34,35 To reduce the crystallization time and crystallite sizes, alternative synthesis methods have been studied, including steam-assisted conversion,36 extensive seeding,34 and post-synthetic methods.37 Tin 3

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 4 of 32

phosphate was prepared for converting glucose into HMF. Tetra-coordinated Sn4+ sites from tin phosphate were identified as the major catalytic species for the isomerization of glucose into fructose.38 Sn(IV) salts are prone to fast and uncontrolled hydrolysis in solution, as noted by the fuming precipitation of tin oxide when SnCl4 is exposed to humid atmospheres.39 Sn species for the preparation of Sn based catalysts should be preferentially present in tetrahedral coordination, and not in octahedral coordination like in SnO2, which showed poor catalytic activities.40 Therefore, synthesis of the Sn based catalyst with high Sn loading without the formation of bulk SnO2 remains a challenge. Porous coordination polymers (PCPs) are formed by coordinating metal ions linked by imidazolate or anionic carboxylate ligands. Due to possessing large surface area, tunable pore sizes, and versatile architectures, PCPs have been developed as promising catalysts.41 Various ligands can be used in preparation of PCPs, thus desired functional groups can be introduced into PCPs. In this work, using 5-sulfoisophthalic acid as the ligand, a coordinated tin-based PCP (SnPCP) has been synthesized on polydopamine-coated MnO2 (MnO2-PDA), as illustrated in Scheme 2. The composite SnPCP@MnO2-PDA contains -SO3H groups and Sn(IV). Thus, SnPCP@MnO2-PDA is a bifunctional composite, which combines the functions of glucose isomerization and fructose dehydration. This novel composite has been used as the catalyst for catalyzing the conversion of glucose into HMF.

4

ACS Paragon Plus Environment

Page 5 of 32 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 Sustainable Chemistry & Engineering

Scheme 2. Schematic presentation for preparation of the composite SnPCP@MnO2-PDA PDA: polydopamine; PCP: porous coordination polymer

METHODOLOGY Materials. The chemicals (chemically pure and analytical reagent grade) were purchased from Sinopharm Chemical Reagent Co. (Shanghai, China) and Sigma−Aldrich (Shanghai, China), they were used without further purification. SnPCP Synthesis. 1.227 g SnCl4·5H2O and 0.939 g 5-sulfoisophthalic acid monosodium salt were added to 30 mL of N,N,-dimethylformamide (DMF), followed by stirring for 15 min. Then the solution was introduced to a teflon-linered stainless steel reactor (50 mL). The reactor was heated at 100°C for 20 h. After cooling to room temperature, the prepared catalysts were filtered through a polycarbonate membrane (average pore size 450 nm), washed with DMF, and then by distilled water, followed by ethanol. The catalysts were dried at vacuum condition at 60°C overnight. Synthesis of MnO2. 2.028 g of MnSO4·H2O, 3.244 g of K2S2O8, and 2.091 g of K2SO4 were dissolved in 300 mL of distilled water. 12.0 mL H2SO4 and 50.96 mg solid AgNO3 were then added into the solution under continuously stirring. The solution was then heated at 45°C for 12 h. The precipitate was collected by filtration, 5

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 6 of 32

washed first with distilled water, followed by ethanol, then dried under vacuum at 60°C for 12 h. Preparation of PDA-coated MnO2 (MnO2-PDA). 50 mg MnO2 was added to Tris-buffer (50 mL, pH 8.5). The mixture was sonicated for 15 min, followed by addition of 100 mg dopamine. Additional sonication (10 min at room temperature) was needed. The MnO2 coating with polydopamine (MnO2-PDA) was filtered. The composite was washed first using distilled water, followed by ethanol. It was dried at vacuum condition at 60°C overnight. Preparation of SnPCP@MnO2-PDA. 1.227 g SnCl4·5H2O and 1.408 g 5-sulfoisophthalic acid monosodium salt were added to 53 mL of DMF, stirred for 15 min. Then 173 mg MnO2-PDA was added to the solution. After stirring for 15 min, the solution was introduced to a Teflon-liner stainless steel reactor (50 mL). The reactor was then heated at 100°C for 20 h. After cooling to room temperature, the prepared catalyst SnPCP@MnO2-PDA was filtered through a polycarbonate membrane (average pore size 450 nm). The catalyst was then washed first with DMF, followed by distilled water, and finally with ethanol. It was then dried at vacuum condition at 60°C for 12 h. Characterization and measurement. The measurement of X-ray photoelectron spectroscopy spectra was performed on a X-ray photoelectron spectrometer (Thermo VG ESCALAB250). The spectra were measured at a pressure (2×10-9 Pa). Mg Ka X-ray was used as the excitation source. Infrared spectra were measured on a FTIR spectrometer (Bruker Tensor 27), which is equipped with a liquid nitrogen-cooled MCT detector, the nominal resolution was 2 cm−1. The N2 sorption was measured on a Micromeritics ASAP 2020 V3.01H at 77 K. Before each measurement, the samples were degassed at 150°C at a high vacuum 6

ACS Paragon Plus Environment

Page 7 of 32 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 Sustainable Chemistry & Engineering

condition (10-5 Torr) for 10 h. Using the Brunauer–Emmett–Teller (BET) method, surface area were calculated based on the adsorption isotherm data. Using the Barret– Joner–Halenda method, the pore size distribution was obtained. The FTIR spectra after pyridine adsorption were studied to evaluate the Lewis acid characteristics on the surface of the SnPCP@MnO2-PDA or SnPCP, which was performed on a Bruker Tensor 27 with a DTGS detector, the resolution was 4 cm-1. Tin cation coordinated in SnPCP@MnO2-PDA and SnPCP was determined by inductively coupled plasma-atomic emission spectroscopy (ICP-AES).

Conversion of glucose to HMF. For catalyzing the conversion of glucose to HMF, 200 mg glucose was dissolved in 5 ml DMSO or other solvent to prepare the glucose solution. Then 50 mg SnPCP@MnO2-PDA or SnPCP was added to the glucose solution. Under strong stirring, the mixture was heated to a desirable temperature using an oil bath. Samples were removed and centrifuged to remove all insoluble particulates. 100 µl of the reaction mixture was diluted with distilled water and taken for analysis. Conversion of fructose to HMF. For catalyzing the conversion of fructose to HMF, 200 mg fructose was dissolved in 5 ml DMSO to prepare the fructose solution. Then 50 mg SnPCP@MnO2-PDA or SnPCP was added to the fructose solution. Under strong stirring, the mixture was heated to a desirable temperature using an oil bath. Samples were removed and centrifuged to remove all insoluble particulates. 50 µl of the reaction mixture was diluted with distilled water and taken for analysis. Substrate and product analysis. Glucose, fructose, formic acid and levulinic acid were analyzed by high-performance liquid chromatography (Shimadzu LC-10A) fitted with a refractive index (RI) detector and an Aminex HPX-87H ion exclusion 7

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 8 of 32

column. The mobile phase was 5 mM H2SO4 flowing at a rate of 0.6 ml/min. The column oven was set to 65°C. The column can efficiently resolute glucose and fructose. HMF was detected using Diamonsil C18 column and an UV–vis detector at 284 nm. The column oven temperature was set at 25°C, and mobile phase was methanol/water = 70:30 (v/v) at a flow rate of 1.0 ml/min. Each reaction was performed in triplicates, and errors were analyzed. Reuse of catalyst. Consecutive use of SnPCP@MnO2-PDA was performed. The catalyst after each run was recovered by filtration, thoroughly washed with water, ethanol, then dried at 60°C under vacuum overnight. The reaction for each cycle was carried out for 1 h.

 RESULTS AND DISCUSSION Characterization of SnPCP and SnPCP@MnO2-PDA. Figure 1 shows the scanning electron microscopy (SEM) images of SnPCP, MnO2, MnO2-PDA, and SnPCP@MnO2-PDA. SnPCP exhibited amorphous nano- to micro-sized particles (Figure 1a). MnO2 itself exhibited microshperes with nanoneedles on their surfaces (Figure 1b). After coating polydopamine (PDA) on MnO2, the nanoneedles cannot be observed, the surface wrinkling is ascribed to the coating of PDA (Figure 1c). The composite SnPCP@MnO2-PDA shows that the amorphous SnPCP particles deposited on MnO2-PDA (Figure 1d). Some SnPCP particles were not on the surface of MnO2-PDA, they were scattered around.

8

ACS Paragon Plus Environment

Page 9 of 32 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 Sustainable Chemistry & Engineering

a

b

c

d Figure 1. SEM images for SnPCP (a), MnO2 (b), MnO2-PDA (c), and SnPCP@MnO2-PDA (d) 9

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 10 of 32

Energy dispersive X-ray spectroscopy (EDS) and XPS spectra analysis. XPS and EDS spectra were recorded for elemental analysis. As shown in Figures 2a, there are significant peaks corresponding to C, O, S and Sn in the EDS image. Figure 2b shows the XPS wide scan spectrum of SnPCP. In the spectrum, the peaks for Sn3d, Sn3p, S2s and S2p appeared. The XPS spectrum for C1s of SnPCP was fitted with Lorentzian and Gaussian lines of variable proportion. XPS spectra were analyzed to reveal the functional moieties on the SnPCP. As shown in supplementary Figure S1a, the peak at 284.7 eV of the C1s XPS spectra was ascribed to C=C/C–C in the aromatic rings. The deconvoluted C1s XPS spectra display two peaks at 286.6 and 288.9, which were assigned to the C-O and carbonyl carbon (C=O) of the ligand (5-sulfoisophthalic acid).42 The S2p peak is split into the S2p1/2 and S2p3/2 peaks (supplementary Figure S1b), they are centered at 168.1 and 169.2 eV, respectively, with a peak area ratio of about 2:1 as expected for the sulfonic acid group. 43

Figure 2. (a) EDS spectrum of SnPCP, (b) wide scan XPS spectra of SnPCP Figure 3a and 3b show the EDS and XPS spectra, respectively, for MnO2, MnO2-PDA and SnPCP@MnO2-PDA. Compared to the EDS and XPS spectra of MnO2, the peaks of C and N appeared in the spectra of MnO2-PDA, confirming the coating MnO2 with PDA. In the EDS spectrum of SnPCP@MnO2-PDA, the peaks for Sn and S appeared. In the XPS spectrum of SnPCP@MnO2-PDA, the peaks for Sn3d, Sn3p, S2s and S2p appeared, confirming the existence of SnPCP. Supplementary 10

ACS Paragon Plus Environment

Page 11 of 32 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 Sustainable Chemistry & Engineering

Figure S2a and S2b show the C1s XPS spectra for MnO2-PDA and SnPCP@MnO2-PDA, respectively. The peak at 284.7 eV was ascribed to C=C/C–C and C-NH2 in the aromatic rings. The two peaks at 288.9 and 286.6 were assigned to and C=O/C=N and C-O/C-N, respectively.42,43 5-Sulfoisophthalic acid was coordinated with Sn in SnPCP@MnO2-PDA, the carboxyl groups are reflected by the shoulder peak at 288.9 (supplementary Figure S2b), this shoulder peak becomes prominent in comparison to that of MnO2-PDA (supplementary Figure S2a). The S2p peak can be splited into S2p1/2 at 168.1 eV and S2p3/2 at 169.2 eV, which were ascribed to the sulfonic acid group (supplementary Figure S2c).44 The EDS and XPS spectra shown in Figure 3 and supplementary Figure S2 confirmed that the Sn coordinated polymer was loaded on MnO2-PDA. The existing states of Sn species for SnPCP@MnO2-PDA and SnPCP samples were investigated by measuring XPS spectra as shown in Figure 4. The peaks at 486.0 and 494.4 eV were ascribed to the Sn atoms in SnO2.45 While the SnPCP@MnO2-PDA and SnPCP samples displayed two peaks of 3d5/2 and 3d3/2 at 487.2 eV and 495.7, respectively.45 The interval of the two peaks is 8.5eV, which indicates the Sn being present in Sn(IV) state.

11

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 12 of 32

Figure 3. Spectra of EDS (a) and wide scan XPS (b) for SnPCP@MnO2-PDA, MnO2-PDA and MnO2. XPS spectra of C1s for MnO2-PDA

12

ACS Paragon Plus Environment

Page 13 of 32 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 Sustainable Chemistry & Engineering

Figure 4. XPS of SnPCP@MnO2-PDA (blue) and SnPCP (black). FTIR spectra. In the MnO2-PDA FTIR spectrum, the band at 1588 cm−1 was assigned to νring (C=C) stretching modes (Figure 5),46 confirming the coating of polydopamine. In the spectra for SnPCP@MnO2-PDA and SnPCP, the band at 1630 cm-1 was assigned to the asymmetric stretching vibrations of C=O. Generally, free carboxyl groups have an absorption band at 1700 cm-1.47 The C=O band was shifted to 1630 cm-1 because of the coordination between Sn4+ and the carboxyl groups in SnPCP@MnO2-PDA and SnPCP. The peaks at 1190 and 1040 cm−1 were associated with symmetric and asymmetric stretching vibrations of the sulfonic acid group.48

13

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 14 of 32

Figure 5. FTIR spectra of SnPCP, SnPCP@MnO2-PDA, and MnO2-PDA.

Pyridine-adsorbed FTIR analysis. Through the FTIR analysis for pyridine adsorption, the Lewis and Brønsted acid sites can be distinced.49 The FTIR spectra of the SnPCP and SnPCP@MnO2-PDA catalysts after pyridine adsorption are shown in Figure 6. For SnPCP, the bands at 1446 and 1604 cm-1 resulted from the pyridine adsorption on Lewis acid sites (Figure 6a), confirming the Lewis acidity of SnPCP. The band at 1558 cm-1 resulted from the C–C stretching vibration of pyridinium ion, which is a characteristic band being used to identify Brønsted acid sites. The peak at 1481 cm−1 was ascribed to the pyridine that had been adsorbed on both Lewis and Brønsted acid sites.13,50 SnPCP@MnO2-PDA exhibited the bands at 1446 and 1604 cm-1 due to the pyridine that had been adsorbed on the Lewis acid sites, at 1558 cm-1 due to the pyridine that had been adsorbed on Brønsted acid sites, and at 1498 cm−1 attributed to pyridine that had been adsorbed on both the Lewis and Brønsted acid sites. All these bands confirmed the Lewis and Brønsted acidities of SnPCP@MnO2-PDA. In 14

ACS Paragon Plus Environment

Page 15 of 32 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 Sustainable Chemistry & Engineering

addition, SnPCP@MnO2-PDA exhibited fewer peaks than SnPCP, indicating the structure difference between the two catalysts. Based on the pyridine FTIR spectra (Figure 6), the Bronsted band area (BBA) was obtained by integrating from 1552 to 1566 cm-1, and the Lewis acid band area (LABA) was obtained by integrating from 1442-1454 cm-1. The ratios of BBA to LABA were 0.16 and 0.91 for SnPCP and SnPCP@MnO2-PDA, respectively. Compared to SnPCP, SnPCP@MnO2-PDA possesses less Lewis acid sites. As the Lewis acid sites are mainly ascribed to the Sn(IV), based on the FTIR spectra it can be deduced that the SnPCP@MnO2-PDA contains less Sn(IV) than SnPCP. The FTIR spectra results are consistent with the ICP-AES elemental analysis, which showed that the amount of Sn(IV) in SnPCP@MnO2-PDA (31.5 wt%) is smaller than that in SnPCP (59.3 wt%). Lewis acidic sites play an important role in the isomerization of glucose to fructose, and Brønsted acidic sites can promote the dehydration of fructose into HMF. A balanced Brønsted/Lewis acids ratio can contribute to achieving higher yield and selectivity for conversion of glucose to HMF.51-53 The Brønsted and Lewis acids sites on SnPCP@MnO2-PDA are better balanced than that on SnPCP.

15

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 16 of 32

Figure 6. Pyridine FTIR spectra for SnPCP and SnPCP@MnO2-PDA.

N2 adsorption and desorption isotherms analysis. The isotherms of SnPCP and SnPCP@MnO2-PDA exhibited a type-IV curve (Figure 7a), suggesting the existence of both micropores and mesopores. The BET specific surface area of SnPCP@MnO2-PDA is 240.6 m2 g-1, which is larger than that of SnPCP (198.9 m2 g-1). SnPCP@MnO2-PDA exhibited a hysteresis for P/P0 from 0.4 to 0.8, while SnPCP exhibited a hysteresis for P/P0 from 0.4 to 0.98, indicating the presence of some larger mesopores in SnPCP. It is also indicated that SnPCP has a larger pore size distribution than SnPCP@MnO2-PDA. The average mesopores diameter for SnPCP@MnO2-PDA is 3.6 nm, which is smaller than that of SnPCP (4.8 nm). The pore size distribution is shown in Figure 7b.

16

ACS Paragon Plus Environment

Page 17 of 32 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 Sustainable Chemistry & Engineering

Figure 7. N2 adsorption and desorption isotherms measured at 77 K for SnPCP and SnPCP@MnO2-PDA (a), pore size and distribution of SnPCP@MnO2-PDA and SnPCP (b) Catalysis using SnPCP@MnO2-PDA and SnPCP. Before catalyzing the conversion of glucose into HMF, the activity of SnPCP and SnPCP@MnO2-PDA for the conversion of fructose into HMF was tested. Figure 8a and 8b show the fructose conversion and HMF yield versus reaction time. After 5 h reaction under the catalysis of SnPCP, 98.2% of fructose was converted and the HMF yield was 92.5% (Figure 17

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 18 of 32

8a). Similar reaction was performed using SnPCP@MnO2-PDA as the catalyst. Figure 8b shows that under the catalysis of SnPCP@MnO2-PDA, 99.6% of fructose was converted, and the HMF yield was 98.0%. The results indicate that the sulfonic groups in the frames of SnPCP and SnPCP@MnO2-PDA are effective for dehydration of fructose into HMF. Then the catalysts SnPCP and SnPCP@MnO2-PDA were used to catalyze the conversion of glucose into HMF. Figure 8c shows the results under the catalysis of SnPCP. After 5 h reaction, 63.7% of glucose was converted and the HMF yield was 39.4%. The HMF selectivity is 61.8%. The yield of fructose was 1.2 % after 5 h. Figure 8d shows results under the catalysis of SnPCP@MnO2-PDA, 92.3% of glucose was converted and the HMF yield was 55.8% after 5 h reaction. The HMF selectivity is 60.5%.

18

ACS Paragon Plus Environment

Page 19 of 32 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 Sustainable Chemistry & Engineering

Figure 8. Fructose conversion and HMF yield versus reaction time under the catalysis of SnPCP (a) and SnPCP@MnO2-PDA(b). Glucose conversion, HMF yield and fructose yield versus reaction time under the catalysis of SnPCP (c) and SnPCP@MnO2-PDA (d). All the reactions were carried out at the same conditions, and the reaction time was 6 h. The results in Figure 8a and 8b shows that both SnPCP and SnPCP@MnO2-PDA are effective for catalyzing the conversion of fructose into HMF. The conversion/yield 19

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 20 of 32

versus time profiles showed that both catalysts exhibited similar activity. The ICP-AES

elemental

analyses

showed

that

the

amounts

of

Sn(IV)

in

SnPCP@MnO2-PDA (31.5 wt%) is smaller than that in SnPCP (59.3 wt%). This means that SnPCP@MnO2-PDA with less loading of Sn(IV) exhibited a similar activity to SnPCP with more loading of Sn(IV). However, for conversion of glucose into HMF, SnPCP@MnO2-PDA showed a higher catalytic activity than SnPCP. To investigate the role of MnO2, PDA, and MnO2-PDA played in the catalysis, they were used as individual catalysts catalyzing the conversion of glucose into HMF. The results in Table 1 show that under the respective catalysis by MnO2 and PDA, fructose and HMF could not be detected. It is indicated that when MnO2 and PDA were used as single catalysts, they did not exhibit any activity for producing fructose and HMF. However, the PDA coated MnO2 (MnO2-PDA) can achieve a fructose yield of 16.3% and a HMF yield of 3.8%. The results in Table 1 suggest that the higher activity achieved by SnPCP@MnO2-PDA is ascribed to the synergistic effect of SnPCP and MnO2-PDA. In addition, Figure 6 shows that the Brønsted and Lewis acids sites on SnPCP@MnO2-PDA are better balanced than that on SnPCP. Figure 7 shows that SnPCP@MnO2-PDA has a smaller average pore size than SnPCP, and has a larger BET specific surface area than SnPCP. The two aspects also contribute to the activity improvement for SnPCP@MnO2-PDA. The TOFs of SnPCP@MnO2-PDA and SnPCP were 2.01 and 0.7 h-1 after the first hour of reaction, respectively, indicating the advantage of SnPCP@MnO2-PDA over SnPCP. Table 1. Activity of MnO2, PDA, and MnO2-PDA for glucose conversion producing fructose and HMF

MnO2 PDA MnO2-PDA

Fructose yield (%)

HMF yield(%)

0 0 16.3%

0 0 3.8%

0.2 g glucose was dissolved in 5 ml DMSO, then 50 mg of MnO2, PDA, and MnO2-PDA each were separately added to the solution. The reaction mixture was heated to 150°C in an oil bath . Reaction 20

ACS Paragon Plus Environment

Page 21 of 32 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 Sustainable Chemistry & Engineering

time was 3 h.

Except DMSO was used as reaction solvent, water, DMA, THF, and the mixture of water/THF were also investigated as solvents under the catalysis of SnPCP@MnO2-PDA. Water is a highly preferred green solvent. DMA is a dipolar aprotic organic solvent and can inhibit the side reactions to some extent.49 HMF can have a good solubility in THF, but glucose has a poor solubility in THF. The results are listed in Table 2. In DMA, water, and THF, the HMF selectivity is comparable. The HMF yields are in sequence of DMA> water > THF. The HMF yield in pure THF is lower than in other solvents due to the poor solubility of glucose has in pure THF. In the mixed solvents water/THF, the HMF selectivity is higher than that in DMA, water, and THF. This is ascribed to the extraction of HMF by THF during the reaction.7 THF is a low boiling point solvent, facilitating subsequent separation of HMF. Figure 9 shows the glucose conversion, fructose yield, and HMF yield in the mixed solvents. After 5 h reaction, SnPCP@MnO2-PDA achieved a HMF yield of 41.2% and a HMF selectivity of 56.3%. Table 2. HMF yield and selectivity in various solvents Solvent

Catalyst

Yield (%)

Selectivity (%)

DMA

SnPCP@MnO2-PDA

41.6

45.2

THF

SnPCP@MnO2-PDA

13.1

46.5

Water

SnPCP@MnO2-PDA

29.1

44.9

Water/THF(1:4)

SnPCP@MnO2-PDA

41.2

56.3

200 mg glucose was dissolved in 5 ml solvent to prepare the glucose solution. 50 mg SnPCP@MnO2-PDA was added to the solution. The reaction temperature was 150 °C. The reaction time was 5 h.

21

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 22 of 32

Figure 9. Glucose conversion, fructose yield and HMF yield versus reaction time under the catalysis of SnPCP@MnO2-PDA in water/THF (v:v 1:4) The catalytic performance of SnPCP@MnO2-PDA in DMSO and biphasic system was compared with other Sn catalyzed reaction systems. As shown in Table 3, the HMF yield and selectivity by SnPCP@MnO2-PDA are higher than that by the homogeneous catalysts SnCl4, Sn phosphate, and Sn-Mont in DMSO.15,38,54 SnPCP@MnO2-PDA gave rise to better results in the mixed solvents water/THF than Sn-Beta in a biphasic system of H2O/1-butanol.7 With the addition of HCl to the reaction solution, the HMF yield and selectivity by Sn-Beta in the mixed solvents H2O/THF/HCl were 56.9% and 72% respectively.7 As HCl alone can contribute to about 10% of the HMF yield, it can be concluded that SnPCP@MnO2-PDA exhibited a comparable catalytic capability with Sn-Beta. Table 3. Comparison of SnPCP@MnO2-PDA with other catalysts HMF yield and selectivity Solvent

Catalyst

Yield

Selec.

T

Time

(%)

(%)

(°C)

(min)

32.7

80

180

15

120

300

38

Ref.

DMSO

SnCl4

10.8

DMSO

Tin phosphate

24.2

DMSO

Sn-Mont

43.6

44.0

160

180

54

DMSO

SnPCP@MnO2-PDA

55.8

60.5

150

300

this work 22

ACS Paragon Plus Environment

Page 23 of 32 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 Sustainable Chemistry & Engineering

H2O/1-butanol

Sn-Beta

13.5

18.0

160

90

7

H2O/THF/HCl

Sn-Beta/HCl

56.9

72.0

180

70

7

H2O/THF

SnPCP@MnO2-PDA

41.2

56.3

150

300

this work

To assess possible deactivation of the catalyst, the reaction for each cycle was controlled for 1 h when the glucose conversion is below 50%.55 The consecutive use of SnPCP@MnO2-PDA was performed in DMSO. Figure 10 shows that after five reuse cycles, the conversion and HMF yield had a very small change, indicating that the catalyst had little deactivation during the recycling catalysis. EDS and XPS spectra were measured for the catalyst SnPCP@MnO2-PDA after reuse. Supplementary Figures S3 and S4 show that the relative intensities of the elements were almost the same before and after reuse of the catalyst, indicating that SnPCP@MnO2-PDA had no significant change after the reuse.

Figure 10. Consecutive use of SnPCP@MnO2-PDA for glucose conversion into HMF. The reaction time was 1 h. The solvent of reaction is DMSO.

 CONCLUSIONS The novel composite SnPCP@MnO2-PDA was synthesized. The Sn-based catalyst was prepared at low temperatures (100 ° C) with no requirement of calcinations. All these contribute to the catalyst preparation in a facile way and not expensive. The composite SnPCP@MnO2-PDA possesses bifunctional catalysis 23

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 24 of 32

ability, catalyzing the tandem reactions of glucose isomerization to fructose and fructose dehydration to HMF. The porous structure of SnPCP facilitates the transport of glucose, intermediate and HMF within the catalyst. In addition, MnO2-PDA was found to be able to catalyze the conversion of glucose to fructose and HMF. The synergistic effect of SnPCP and MnO2-PDA achieved HMF yields of 55.8% in DMSO and 41.2% in the mixed solvents water/THF. The Brønsted and Lewis acids sites are appropriately balanced on SnPCP@MnO2-PDA, contributing to achieving high HMF yield and selectivity. Consecutive use of SnPCP@MnO2-PDA demonstrated that after 5 cycles, the activity loss is not significant in terms of the HMF yield and glucose conversion.

■ ASSOCIATED CONTENT Supporting Information There are two pages. Figure S1 for the XPS spectra of SnPCP, C1s of SnPCP, S2p of SnPCP. Figure S2 for the XPS spectra of C1s for MnO2-PDA and SnPCP@MnO2-PDA, and XPS spectra of S2p for SnPCP@MnO2-PDA. Figure S3 for the EDS analysis for the catalyst SnPCP@MnO2-PDA after reuse. Figure S4 for the XPS analysis for the catalyst SnPCP@MnO2-PDA after reuse



AUTHOR INFORMATION

Corresponding Authors *Tel.: +86-10-64423254. Author Contributions K. Li and M. Du contributed equally to this work Notes The authors declare no competing financial interest.



ACKNOWLEDGEMENTS

This work was supported by the National Science Foundation of China (21476023).

24

ACS Paragon Plus Environment

Page 25 of 32 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 Sustainable Chemistry & Engineering



REFERENCES (1) Cui, M. S.; Deng, J.; Li, X. L.; Fu Y. Production of

4‑hydroxymethylfurfural from derivatives of biomass-derived glycerol for chemicals and polymers. ACS Sustainable Chem. Eng. 2016, 4, 1707−1714. DOI:10.1039/B611568C.

(2) Arias, K. S.; Climent, M. J.; Corma, A.; Iborra, S. Two-dimensional itq‑2 zeolite for biomass transformation: synthesis of alkyl 5‑benzyl-2furoates as intermediates for fine chemicals. ACS Sustainable Chem. Eng. 2016, 4, 6152−6159. DOI: 10.1021/acssuschemeng.6b01662 (3) Yao, S.; Wang, X.; Jiang, Y.; Wu, F.; Chen, X.; Mu, X. One-step conversion of biomass-derived 5‑hydroxymethylfurfural to 1,2,6-hexanetriol over Ni−Co−Al mixed oxide catalysts under mild conditions. ACS Sustainable Chem. Eng. 2014, 2, 173−180. DOI: 10.1021/sc4003714 (4) Yang, Q.; Sherbahn, M.; Runge T. Basic amino acids as green catalysts for isomerization of glucose to fructose in water. ACS Sustainable Chem. Eng. 2016, 4, 3526−3534. DOI: 10.1021/acssuschemeng.6b00587 (5) Moreno-Recio, M.; Santamaría-González, J.; Maireles-Torres, P. Brönsted and Lewis acid ZSM-5 zeolites for the catalytic dehydration of glucose into 5-hydroxymethylfural. Chem. Eng. J. 2016, 303, 22-30. DOI:10.1016/ j.cej.2016.05.120 (6) Tang, J.; Guo, X.; Zhu, L.; Hu, C. Mechanistic study of glucose-to-fructose isomerization in water catalyzed by [Al(OH)2(aq)]+. ACS Catal. 2015, 5, 5097-5103. DOI: 10.1021/acscatal.5b01237 (7) Nikolla, E.; Román-Leshkov, Y.; Moliner, M.; Davis, M. E. “One-Pot” synthesis of 5-(hydroxymethyl)-furfural from carbohydrates using Tin-beta zeolite, ACS Catal. 2011, 1, 408-410. DOI: 10.1021/cs2000544. (8) Yu, S.; Kim, E.; Park, S.; Song, I. K.; Jung, C. Isomerization of glucose into fructose over Mg-Al hydrotalcite catalysts. Catal. Comm. 2012, 29, 63-67. DOI: 10.1016/j.catcom.2012.09.015 (9) Yue, C.; Magusin, P. C. M. M.; Mezari, B.; Rigutto, M.; Hensen, E. J. M. Hydrothermal synthesis and characterization of a layered zirconium silicate. Microporous Mesoporous Mater. 2013, 180, 48-55. DOI: 10.1016/ 25

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 26 of 32

j.micromeso.2013.06.032. (10) Saravanamurugan, S.; Paniagua, M.; Melero, J. A.; Riisager, A. Efficient isomerization of glucose to fructose over zeolites in consecutive reactions in alcohol and aqueous media. J. Am. Chem. Soc. 2013, 135, 5246-5249. DOI: 10.1021/ja400097f. (11) Liu, C.; Carraher, J. M.; Swedberg, J. L.; Herndon, C. R.; Fleitman, C. N.; Tessonnier, J. Selective base-catalyzed isomerization of glucose to fructose. ACS Catal. 2014, 4, 4295-4298. DOI: 10.1021/cs501197w. (12) Chakrabarti, A.; Sharma, M. M. Cationic ion exchange resins as catalyst.

React.

Polym.

1993,

20,

1-45.

DOI: 10.1016/0923-1137(93)90064-M (13) Atanda, L.; Mukundan, S.; Shortri, A.; Ma, Q.; Beltramini, J. Catalytic conversion of glucose to 5-hydroxymethylfurfural with a phosphated TiO2 catalyst. Chem. Cat. Chem. 2015, 7, 781-790. DOI: 10.1002/cctc.201402794. (14) Guo, T.; Tong, X. L.; Cheng, Y.; Xue, S. Tin-catalyzed efficient conversion of carbohydrates for the production of 5-hydroxymethylfurfural in the presence of quaternary ammonium salts. Carbohydr. Res. 2013, 370, 33-37. DOI: 10.1016/j.carres.2013.01.012. (15) Hu, S.; Zhang, Z.; Song, J.; Zhou, Y.; Han, B. Efficient conversion of glucose into 5-hydroxymethylfurfural catalyzed by a common Lewis acid SnCl4 in an ionic liquid. Green Chem. 2009, 11, 1746-1749. DOI: 10.1039/b914601f. (16) Moliner, M.; Román-Leshkov, Y.; Davis, M. E. Tin-containing zeolites are highly active catalysts for the isomerization of glucose in water. PNAS 2010, 107, 6164-6168. DOI: 10.1073/pnas.1002358107. (17) Deval, R. B.; Assary, R. S.; Nikolla, E.; Moliner, M.; Román-Leshkov, Y.; Hwang, S.; Palsdottir, A.; Silverman, D.; Lobo, R. F.; Curtiss, L. A.; Davis, M. E. Metalloenzyme-like catalyzed isomerizations of sugars by Lewis

acid

zeolites.

PNAS

2012,

109,

9727-9732.

DOI:

10.1073/pnas.1206708109. (18) Chang, C.; Cho, H. J.; Wang, Z.; Wang, X.; Fan, W. Fluoride-free synthesis of a Sn-BEA catalyst by dry gel conversion. Green Chem. 2015, 17, 2943-2951. DOI: 10.1039/C4GC02457E. 26

ACS Paragon Plus Environment

Page 27 of 32 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 Sustainable Chemistry & Engineering

(19) Jin, J.; Ye, X.; Li, Y.; Wang, Y.; Li, L.; Gu, J.; Zhao, W.; Shi, J. Synthesis of mesoporous beta and Sn-Beta zeolites and their catalytic performances. Dalton Trans. 2014, 43, 8196-8204. DOI: 10.1039/ c4dt00567h. (20) Liu, M.; Jia, S.; Li, C.; Zhang, A.; Song, C.; Guo, X. Facile preparation of Sn-β zeolites by post-synthesis (isomorphous substitution) method for isomerization of glucose to fructose. Chin. J. of Catal. 2014, 35, 723-732. DOI: 10.1016/S1872-2067(14)60071-1. (21) Padovan, D.; Parsons, C.; Grasina, M. S.; Hammond, C. Intensification and deactivation of Sn-beta investigated in the continuous regime. Green Chem. 2016, 18, 5041–5049. DOI: 10.1039/C6GC01288D. (22) Koehle, M.; Lobo, R. F. Lewis acidic zeolite Beta catalyst for the Meerwein–Ponndorf–Verley reduction of furfural. Catal. Sci. Technol. 2016, 6, 3018–3026. DOI: 10.1039/C5CY01501D. (23) Tolborg, S.; Katerinopoulou, A.; Falcone, D. D.; S´adaba, I.; Osmundsen, C. M.; Davis, R. J.; Taarning, E.; Fristrup, P.; Holm, M. S. Incorporation of tin affects crystallization,morphology, and crystal composition of Sn-Beta. J. Mater. Chem. A 2014, 2, 20252–20262. DOI: 10.1039/C4TA05119J. (24) Bermejo-Deval, R.; Gounder, R.; Davis, M. E. Framework and extraframework Tin sites in zeolite Beta react glucose differently. ACS Catal. 2012, 2, 2705−2713. DOI: 10.1021/cs300474x. (25) Tian, X.; Jiang, Z.; Jiang, Y.; Xu, W.; Li, C.; Luo, L.; Jiang, Z. Sulfonic acid-functionalized mesoporous carbon/silica as efficient catalyst for dehydration of fructose into 5-hydroxymethylfurfural. RSC Adv. 2016, 6, 101526–101534. DOI: 10.1039/C6RA20304C. (26) Shi, C.; Xin, J.; Liu, X.;

Lu, X.; Zhang, S. Using sub/ supercritical

CO2 as “phase separation switch” for the efficient production of 5-hydroxymethylfurfural from fructose in an ionic liquid/organic biphasic system. ACS Sustainable Chem. Eng. 2016, 4 (2), 557–563. DOI: 10.1021/acssuschemeng.5b00889. (27)

Eminov, S.; Wilton-Ely, J. D. E. T.; Hallett J. P. Highly selective and 27

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 28 of 32

near-quantitative conversion of fructose to 5-hydroxymethylfurfural using mildly acidic ionic liquids. ACS Sustainable Chem. Eng., 2014, 2 (4), 978– 981. DOI: 10.1021/sc400553q. (28) Vekariya, R. H.; Patel, H. D. Sulfonated polyethylene glycol (PEG-OSO3H) as a polymer supported biodegradable and recyclable catalyst in green organic synthesis: recent advances. RSC Adv. 2015, 5, 49006– 49030. DOI: 10.1080/15533174.2014.900799. (29) Hou, Q.; Li, W.; Ju, M.; Liu, L.; Chen, Y.; Yang, Q. One-pot synthesis of sulfonated graphene oxide for efficient conversion of fructose into HMF. RSC Adv. 2016, 6, 104016–104024. DOI:10.1039/C6RA23420H. (30) Simeonov, S. P.; Coelho, J. A. S.; Afonso, C. A. M. Integrated chemo-enzymatic production of 5-hydroxymethylfurfural from glucose. ChemSusChem 2013, 6, 997–1000. DOI: 10.1002/cssc.201300176. (31) Swift, T. D.; Nguyen, H.; Anderko, A.; Nikolakis, V.; Vlachos, D. G. Tandem Lewis/Brønsted homogeneous acid catalysis: conversion of glucose to 5-hydoxymethylfurfural in an aqueous chromium(III) chloride and hydrochloric acid solution. Green Chem. 2015, 17, 4725–4735. DOI: 10.1039/C5GC01257K. (32) Wang, X.; Zhang, H.; Ma, J.; Ma, Z. Bifunctional Brønsted–Lewis solid acid as a recyclable catalyst for conversion of glucose to 5-hydroxymethylfurfural and its hydrophobicity effect. RSC Adv. 2016, 6, 43152–43158. DOI:10.1039/C6RA03565E. (33)

Herbst,

A.;

Janiak,

C.

Selective

glucose

conversion

to

5-hydroxymethylfurfural (5-HMF) instead of levulinic acid with MIL-101Cr MOF-derivatives.

New

J.

Chem.

2016,

40,

7958—7967.

DOI:

10.1039/c6nj01399f. (34) Chang, C.; Wang, Z.; Dornath, P.; Cho, H. J.; Fan, W. Rapid synthesis of Sn-Beta for the isomerization of cellulosic sugars. RSC Adv. 2012, 2, 10475–10477. DOI:10.1039/C2RA21381H. (35)

Ferrini, P.; Dijkmans, J.; Clercq, R. D.; Van de Vyver, S.; Dusselier,

M.; Jacobs, P. A.; Sels, B. F. Lewis acid catalysis on single site Sn centers incorporated into silica hosts. Coord. Chem. Rev. 2017, 343, 220–255. 28

ACS Paragon Plus Environment

Page 29 of 32 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 Sustainable Chemistry & Engineering

DOI:10.1016/j.ccr.2017.05.010. (36) Kang, Z.; Zhang, X.; Liu, H.; Qiu, J.; Yeung. K. L. A rapid synthesis route for Sn-Beta zeolites by steam-assisted conversion and their catalytic performance in Baeyer–Villiger oxidation. Chem. Eng. J. 2013, 218, 425– 432. DOI:10.1016/j.cej.2012.12.019. (37) Hammond, C.; Conrad, S.; Hermans, I. Simple and scalable preparation of highly active Lewis acidic Sn-β. Angew. Chem. Int. Ed. 2012, 51, 11736 –11739. DOI: 10.1002/anie.201206193. (38) Hou, Q.; Zhen, M.; Liu, L.; Chen, Y.; Huang, F.; Zhang, S.; Li, W.; Ju, M. Tin phosphate as a heterogeneous catalyst for efficient dehydration of glucose into 5-hydroxymethylfurfural in ionic liquid. Appl. Catal. B: Environ. 2018, 224 , 183–193. DOI: 10.1016/j.apcatb.2017.09.049. (39) Neilson, J. R.; Kurzman, J. A.; Seshadri, R.; Morse, D. E. Ordering double perovskite hydroxides by kinetically controlled aqueous hydrolysis. Inorg. Chem. 2011, 50, 3003–3009. DOI: 10.1021/ic1025075. (40) Dijkmans, J.; Dusselier, M.; Gabriëls, D.; Houthoofd, K.; Magusin, P. C. M. M.; Huang, S.; Pontikes, Y.; Trekels, M.; Vantomme, A.; Giebeler, L.; Oswald, S.; Sels, B. F. Cooperative catalysis for multistep biomass conversion with Sn/Al Beta zeolite. ACS Catal. 2015, 5, 928−940. DOI: 10.1021/cs501388e. (41) Huang, Y.; Liang, J.; Wang, X.; Cao, R. Multifunctional metal–organic framework catalysts: synergistic catalysis and tandem reactions. Chem. Soc. Rev. 2017, 46, 126-157. DOI:10.1039/C6CS00250A. (42) Majeed, I.; Nadeem, M. A.; Badshah, A.; Kanodarwala, F. K.; Ali, H.; Khan, M. A.; Stride, J. A.; Nadeem, M. A. Titania supported MOF-199 derived

Cu–Cu2O

nanoparticles:

highly

efficient

non-noble

metal

photocatalysts for hydrogen production from alcohol–water mixtures. Catal. Sci. Technol. 2017, 7, 677–686. DOI:10.1039/C6CY02328B. (43) Hwang, D. S.; Masic, A.; Prajatelistia, E.; Iordachescu, M.; Waite, J. H. Marine hydroid perisarc: A chitin- and melanin-reinforced composite with DOPA–iron(III) complexes. Acta Biomaterialia 2013, 9, 8110–8117. 29

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 30 of 32

DOI:10.1016/j.actbio.2013.06.015. (44) Gui, A. L.; Liu, G.; Chockalingam, M.; Saux, G. L.; Harper, J. B.; Gooding, J. J. A comparative study of modifying gold and carbon electrode with 4-sulfophenyl diazonium salt. Electroanalysis 2010, 22, 1283 – 1289. DOI: 10.1002/elan.200900615. (45) Dai W.; Wang C.; Tang B.; Wu G.; Guan N.; Xie Z.; Hunger M.; Li L. Lewis acid catalysis confined in zeolite cages as a strategy for sustainable heterogeneous hydration of epoxides. ACS Catal. 2016, 6, 2955−2964. DOI: 10.1021/acscatal.5b02823. (46) Zangmeister, R. A.; Morris, T. A; Tarlov, M. J. Characterization of polydopamine thin films deposited at short times by autoxidation of dopamine. Langmuir 2013, 29, 8619-8628. DOI: 10.1021/la400587j. (47) Weng, H.; Yan, B. Cadmium metal–organic frameworks: Ln3+ ion functionalized assembly, fluorescence tuning and polymer film preparation. New J. Chem. 2016, 40, 3732-3737. DOI:10.1039/C5NJ02999F. (48) Kurc, T.; Janczak, J.; Hoffmann, J.; Videnova-Adrabinska, V. New heterometallic

hybrid

polymers

constructed

with

aromatic

sulfonate-carboxylate ligands: synthesis, layered structures, and properties. Cryst. Growth Des. 2012, 12, 2613−2624. DOI: 10.1021/cg300199j. (49) Emeis, C. A. Determination of integrated molar extinction coefficients for infrared absorption bands of pyridine adsorbed on solid acid catalysts. J. catal. 1993, 141, 347-354. DOI: 10.1006/jcat.1993.1145. (50) Nakajima, K.; Noma, R.; Kitano, M.; Hara, M. Selective glucose transformation by titania as a heterogeneous Lewis acid catalyst. J. Mol. Catal.

A:

Chem.

2014,

388-389,

100-105.

DOI:

10.1016/j.molcata.2013.09.012. (51) Xue, Z.; Ma, M.; Li, Z.; Mu, T. Advances in the conversion of glucose and cellulose to 5-hydroxymethylfurfural over heterogeneous catalysts. RSC Adv. 2016, 6, 98874–98892. DOI:10.1039/C6RA20547J. (52) Ordomsky, V. V.; van der Schaaf, J.; Schouten, J. C.; Nijhuis, T. A. Glucose dehydration to 5-hydroxymethylfurfural in a biphasic system over solid

acid

foams.

ChemSusChem

2013,

6,

1697–1707.

DOI:

10.1002/cssc.201300017. 30

ACS Paragon Plus Environment

Page 31 of 32 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 Sustainable Chemistry & Engineering

(53) Li, X.; Xia, Q.; Nguyen, V. C.; Peng, K.; Liu, X.; Essayem, N.; Wang, Y. High yield production of HMF from carbohydrates over silica–alumina composite

catalysts.

Catal.

Sci.

Technol.

6,

2016,

7586–7596.

DOI:10.1039/C6CY01628F. (54) Wang, J.; Ren, J.; Liu, X.; Xi, J.; Xia, Q.; Zu, Y.; Lu G.; Wang, Y. Direct conversion of carbohydrates to 5-hydroxymethylfurfural using Sn-Mont

catalyst.

Green

Chem.

14,

2012,

2506–2512.

DOI:10.1039/C2GC35699F. (55) Lari, G. M.; Dapsens, P. Y.; Scholz, D.; Mitchell, S.; Mondelli, C.; Pérez-Ramírez, J. Deactivation mechanisms of tin-zeolites in biomass conversions.

Green

Chem.

2016,

18,

1249–1260.

DOI:10.1039/C5GC02147B.

31

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

Page 32 of 32

For Table of Contents Use Only

The novel composite SnPCP@MnO2-PDA, which was synthesized in a facile way, possesses bifunctional catalysis ability for producing HMF from glucose.

32

ACS Paragon Plus Environment