Next Article in Journal
C-O Bond Hydrogenolysis of Aqueous Mixtures of Sugar Polyols and Sugars over ReOx-Rh/ZrO2 Catalyst: Application to an Hemicelluloses Extracted Liquor
Previous Article in Journal
Artificial Intelligence Modelling Approach for the Prediction of CO-Rich Hydrogen Production Rate from Methane Dry Reforming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sn-Based Porous Coordination Polymer Synthesized with Two Ligands for Tandem Catalysis Producing 5-Hydroxymethylfurfural

Department of Chemical Engineering, Beijing University of Chemical Technology, Beijing 100029, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(9), 739; https://doi.org/10.3390/catal9090739
Submission received: 14 July 2019 / Revised: 9 August 2019 / Accepted: 21 August 2019 / Published: 31 August 2019
(This article belongs to the Section Catalytic Materials)

Abstract

:
5-Hydroxymethylfurfural (HMF) is a biomass-derived important platform compound. Developing an efficient catalyst for producing HMF from a biomass source is important. Herein, using the ligands 5-sulfoisophthalic acid (SPA) and imidazole (Imd), a tin-based porous coordination polymer was synthesized, namely SPA-Imd-TinPCP. This novel material possesses a multifunctional catalysis capability. The coordinated tin (IV) can catalyze the isomerization of glucose to fructose. The ligand imidazole, as an additional base site, can catalyze glucose isomerization. The sulfonic group of the ligand SPA can catalyze the dehydration of fructose to HMF. SPA-Imd-TinPCP was used as a catalyst for the conversion of glucose to HMF. HMF yields of 59.5% in dimethyl sulfoxide (DMSO) and 49.8% in the biphasic solvent of water/tetrahydrofuran were obtained. Consecutive use of SPA-Imd-TinPCP demonstrated that, after reusing it five times, there was no significant activity loss in terms of the glucose conversion and HMF yield.

1. Introduction

5-Hydroxymethylfurfural (HMF) is an important platform chemical. It can be used as a starting material for the synthesis of value-added chemicals [1,2]. The glucose conversion to HMF is an attractive approach, as glucose can be obtained from the biomass resource [3]. Scheme 1 presents the two steps of the approach. Glucose is first isomerized to fructose, followed by the dehydration of fructose to HMF [4,5].
Developing efficient catalysts is essential to make the approach economically competitive. Homogeneous and heterogeneous chemical catalysts have been investigated for the isomerization of glucose to fructose [6,7,8,9,10,11,12,13]. Glucose isomerization to fructose was catalyzed by both Brønsted bases and Lewis acids. However, monosaccharides are unstable under strong alkaline conditions. Organic bases, including basic amino acid, acyclic and cyclic amines, have been investigated for glucose isomerization to fructose [7,8]. Lot of research has also focused on the Lewis acid-catalyzed route for glucose isomerization [10,11,12,13,14]. Sn-Beta and Lewis acidic metal salt, SnCl4, are promising catalysts for glucose isomerization [10,11,12,13,14]. Sn-Beta is an effective glucose isomerization catalyst, but it is not commercially available yet, and its synthesis is complex [15,16]. Sn(IV) salts are likely to hydrolyze in a fast and uncontrolled manner. For example, when SnCl4 is exposed to a humid atmosphere, the fuming precipitation of tin oxide can be observed [17]. For the preparation of Sn-based catalysts, Sn species should be tetrahedrally coordinated, not octahedrally coordinated like in SnO2, which exhibited poor activity [18]. Therefore, for synthesis of Sn-based catalysts with a high Sn loading, avoiding the formation of bulk SnO2 should be considered.
Liquid acids, such as H2SO4 and H3PO4, are efficient catalysts for the dehydration of fructose into HMF. Using homogeneous catalysts can lead to environmental pollution and increase the separation costs [19]. Solid acid catalysts, like sulfonate groups or phosphate-group-functionalized solid acids, are highly desirable from an economic and environmental point of view [19,20,21]. Catalysts possessing both Lewis/Brønsted acids sites exhibit bifunctional properties, and can be used for the one-pot synthesis of HMF from glucose [22,23,24]. Separation and purification of the intermediates is not necessary in the one-pot approach. Tin phosphate was prepared and used for converting glucose into HMF, and an HMF yield of 58.3% in the ionic liquid EMIMBr was obtained [22]. A polydivinylbenzene-based catalyst was used for the conversion of glucose to HMF, and HMF was obtained with stable yields [23]. Sulfonated MIL-101Cr achieved a 29% glucose conversion for producing HMF after a 24 h reaction in H2O/tetrahydrofuran (THF) [1].
Porous coordination polymers (PCPs) are formed by coordinating metal ions with organic ligands. Because of their tunable pore sizes, large surface area, and versatile architectures, PCPs have been investigated as candidates for various applications [25]. 5-Sulfoisophthalic acid, with two carboxyl groups and a sulfonic group, can be used as a ligand to synthesize PCPs [26,27,28]. Imidazole is widely used as a ligand, coordinated with transition metals forming PCPs [29,30]. Herein, 5-sulfoisophthalic acid (SPA) was used as the ligand and imidazole was used as an additional ligand; these were coordinated with tin (IV) to synthesize the tin-based PCP (SPA-Imd-TinPCP). SPA-Imd-TinPCP exhibited bifunctional catalysis capability. With tin (IV) as the Lewis acid sites and imidazole as the Brønsted base sites, SPA-Imd-TinPCP can catalyze the glucose isomerization reaction. The uncoordinated sulfonate groups of SPA on SPA-Imd-TinPCP can catalyze the dehydration of fructose to HMF. Thus, SPA-Imd-TinPCP combines the functions of isomerization of glucose into fructose and dehydration of fructose into HMF. This novel catalyst has been used and evaluated for catalyzing the conversion of glucose into HMF.

2. Results and Discussion

2.1. Characterization of SPA-Imd-TinPCP

A transmission electron microscope (TEM) image (Figure 1a) shows that SPA-Imd-TinPCP is a polymorph made up of partial crystals. The X-ray diffraction (XRD) pattern (Figure 1b) shows intense broad diffraction peaks corresponding to a polymorph made up of partial crystals. The X-ray diffraction (XRD) pattern is in accordance with the TEM image. Energy-disperse spectroscopy (EDS) and X-ray photoelectron spectroscopy (XPS) were used for the analysis of elements. The spectra of EDS (Figure 2a) and XPS (Figure 2b) show that peaks corresponding to the elements C, O, S, N, and Sn appeared. The appearance of the peaks of S and N is ascribed to the coordinated ligands of SPA and imidazole.
The XPS spectrum for N 1s of SPA-Imd-TinPCP was fitted with Lorentzian and Gaussian lines of variable proportion. As shown in Figure 3, the two peaks at 401.4 and 400.0 eV were assigned to the binding energies for C‒NH‒C [30] and chemisorbed imidazole [31], respectively. Imidazole contains two nitrogen atoms, a pyrrole nitrogen (N-1) and a pyrrodine nitrogen (N-3). The unshared pair of electrons on N-3 usually coordinated with metal ions via the unshared pair of electrons [31]. The spectrum decomposition confirmed that the ligand imidazole was coordinated with tin (IV) in the structure of SPA-Imd-TinPCP. Figure 4a shows the spectrum decomposition for C1s. The three peaks at 284.7, 286.6, and 289.0 eV were assigned to C‒C/C‒H, C‒O, and O‒C=O [32], respectively. These peaks were ascribed to the presence of the ligand 5-sulfoisophthalic acid. The peaks at 285.2 and 286.0 eV was assigned to C=N and C‒N [33], which were attributed to the presence of imidazole. Figure 4b shows the existing state of tin species in SPA-Imd-TinPCP. The two peaks for Sn 3d5/2 and Sn 3d3/2 appeared at 487.2 and 495.7 eV, respectively. They appeared at positions different from those for the tin atoms in SnO2 (486.0 and 494.4 eV) [34], indicating the presence of tin (IV) in SPA-Imd-TinPCP.
Figure 5 shows the Fourier transform infrared (FTIR) spectrum for SPA-Imd-TinPCP. The band at 1630 cm−1 was assigned to the asymmetric stretching vibrations of C=O. Free carboxyl groups have an absorption band at 1700 cm−1 [35]. The wavenumber was shifted to 1630 cm−1, indicating that the carboxyl group of SPA was coordinated with tin (IV). The peaks at 1190 and 1045 cm−1 were ascribed to the symmetric and asymmetric stretching vibrations of the sulfonic acid group [36]. The peak at 1322 cm−1 was assigned to the ring C‒N symmetric stretching of imidazole [31,37]. The band at 1438 cm−1 was ascribed to the bending C=N vibration of the imidazole ring [38].
Figure 6 shows the FTIR spectrum with pyridine adsorption for the analysis of Brønsted and Lewis acid sites. The band at 1596 cm−1 is associated with pyridine interact with the catalyst with a hydrogen bond mode [39]. The band at 1445 cm−1 is attributed to the pyridine adsorption on Lewis acid sites [39,40,41,42]. The appearance of these bands confirmed that SPA-Imd-TinPCP exhibited a Lewis acidity. The bands at 1540 and 1640 cm−1 are ascribed to the C-C stretching vibration of pyridinium ion, indicating the presence of Brønsted acid sites. The peak at 1488 cm−1 was attributed to pyridine, adsorbed on both Lewis and Brønsted acid sites [39,40,41,42].
Figure 7a shows the N2 adsorption and desorption isotherms measured at 77 K, and Figure 7b shows the pore size distribution. SPA-Imd-TinPCP exhibited a large Brunauer-Emmett-Teller (BET) surface area of 345.2 m2 g−1. The uptake under the lower pressure range ( P / P 0 = 10 5 to 0.1 ) demonstrated a microporous feature. The uptake occurred near P / P 0 = 0.9 ~ 1.0 can be attributed to the textural pores created by the aggregation of nanocrystals. Most of the pore diameters of SPA-Imd-TinPCP are smaller than 4 nm.

2.2. The Roles 5-Sulfoisophthalic Acid and Imidazole Play in Catalysis

2.2.1. 5-Sulfoisophthalic Acid

The ligand sulfoisophthalic acid has one sulfonic acid group and two carboxyl groups. In comparison to the sulfonic acid group, the carboxyl groups have a much stronger coordination ability. Thus SPA-Imd-TinPCP contains sulfonic acid groups that are not coordinated with Tin (IV), as illustrated in Figure 8. These sulfonic acid groups are available to catalyze the dehydration of fructose to HMF. Figure 9 shows the profile of fructose conversion versus reaction time under the catalysis of SPA-Imd-TinPCP. After a 3-h reaction, the fructose conversions were 92.9% in a biphasic solvent of H2O/THF (v/v 1:4) and 99.9% in DMSO. The HMF yields were 72.9% in H2O/THF (v/v 1:4) and 99.8% in DMSO. The results demonstrated that the sulfonic acid group of SPA-Imd-TinPCP exhibited a high efficiency for the dehydration of fructose into HMF.

2.2.2. Imidazole

Imidazole is another ligand in SPA-Imd-TinPCP. When preparing SPA-Imd-TinPCP, the initial molar ratio of 5-sulfoisophthalic acid to imidazole was 2:1. Imidazole comprises two nitrogen atoms, pyrrole nitrogen (N-1) and pyrrodine nitrogen (N-3). The N1s spectra (Figure 3) confirmed that the nitrogen atom N-3 was coordinated with tin (IV). We speculate on the coordination of imidazole with tin (IV) in the form as illustrated in Figure 8. To investigate the role imidazole plays in catalysis, we used it to catalyze the glucose isomerization. Imidazole achieved a glucose conversion of 63.2%, a fructose yield of 21.1%, and an HMF yield of 1.9% (Table 1). SPA-TinPCP is a Sn-based porous coordination polymer, which was synthesized only using 5-sulfoisophthalic acid as the ligand. SPA-TinPCP achieved a glucose conversion of 63.2% and an HMF yield of 37.3%. In contrast, SPA-Imd-TinPCP achieved a glucose conversion of 86.2% and an HMF yield of 49.8%. The ICP-AES elemental analysis showed that the amount of tin (IV) in SPA-Imd-TinPCP (39.5 wt %) is comparable to that of tin (IV) in SPA-TinPCP (38.7 wt %). It can be reasoned that the ligand imidazole made a contribution to the better catalysis efficiency of SPA-Imd-TinPCP, as presented in Table 1.
Reaction conditions: 200 mg of glucose was dissolved in 5 mL of H2O/THF (1:4) to prepare the glucose solution, and 50 mg of catalyst was added to the solution. The reaction temperature and pressure were 160 °C and 5 atm, respectively. The reaction time was 5 h.

2.3. Catalysis of Conversion of Glucose to HMF

Two reaction media were used for the conversion of glucose to HMF under the catalysis of SPA-Imd-TinPCP. Figure 10a shows the results in DMSO. After a 4 h reaction, 89.5% of glucose was converted, the HMF yield was 59.5%, equivalent to an HMF selectivity of 66.5%, and the yield of fructose was 1.0%. Figure 10b shows the results in a biphasic solvent of H2O/THF (v/v 1:4). After a 5-h reaction, 86.2% of glucose was converted and the HMF yield was 49.8%, equivalent to an HMF selectivity of 57.8%.
After a 5 h reaction in H2O/THF (v/v 1:4), separation and purification of HMF from the reaction mixture were complete. The insoluble byproducts and catalysts were removed from the reaction mixture by filtration. The filter cake was washed with deionized water. After adding saturated sodium chloride into the filtrate of the reaction mixture, the mixture (about 10 mL) was separated into two phases: the upper layer of THF and the lower aqueous phase. The concentration of HMF in both phases were analyzed with high performance liquid chromatography (HPLC). It was found that about one-fourth of the HMF produced still existed in the aqueous layer. In order to extract HMF in the aqueous phase into organic phase, 15 mL of diethyl ether was added into the filtrate as an additional extraction solvent. With the addition of diethyl ether, about 92% of HMF was extracted into the organic layer. The THF and diethyl ether were separated by distillation at 50 °C. Pure HMF was obtained as the residue in the flask.
Consecutive use of SPA-Imd-TinPCP was performed with a 5-h reaction for each cycle in DMSO and in H2O/THF. The HMF yield and glucose conversion showed that after reusing it five times, the activity loss of SPA-Imd-TinPCP can be neglected (Figure 11). SPA-Imd-TinPCP has been compared with other tin-based catalysts. For the catalysis in DMSO, SPA-Imd-TinPCP showed advantages over the homogeneous catalysts SnCl4, tin phosphate, and Sn-Mont in terms of the HMF yield and selectivity, as presented in Table 2. In the biphasic solvent H2O/THF, SPA-Imd-TinPCP exhibited better HMF yield and selectivity than that of Sn-Beta in water/1-butanol. Sn-Beta in H2O/THF/HCl exhibited an HMF yield of 56.9% and an HMF selectivity of 72% [5]. Because HCl alone can achieve an HMF yield of about 10%, it can be reasoned that SPA-Imd-TinPCP had comparable results to Sn-Beta in H2O/THF.
Other metal-based catalysts have been developed for the catalytic conversion of glucose to HMF with high efficiency. TiO2-ZrO2 modified with CH3COOH achieved the highest yield of 76% to HMF in H2O/THF [44]. The synthesis of modified TiO2-ZrO2 [44] was carried out with the sol-gel method. The gel was subjected to the extraction of liquids in an oven at 70 °C for 48 h and then calcined at 500 °C for 12 h. The HMF yields of modified TiO2-ZrO2 were reduced by about 60% after three cycles due to the loss of acidic sites. A bifunctional porous coordination polymer, PCP(Cr)-NH2, achieved an HMF yield of 65.9% from glucose in H2O/THF [45]. However, PCP(Cr)-NH2 was synthesized at a higher temperature, 220 °C, and hydrogen fluoride was used for preparing the catalyst [45]. Compared with the reported catalyst, SPA-Imd-TinPCP showed advantages including catalytic stability, a lower synthesis temperature (100 °C), a synthesis process free of hydrofluoric acid, and calcinations not being needed.

3. Experimental Section

3.1. Chemicals and Materials

All chemicals required for the synthesis of catalyst and catalytic reactions were purchased from Sigma-Aldrich (Shanghai, China) or Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) and used as received without further purification.

3.2. Synthesis of SPA-Imd-TinPCP

A 2.454 g portion of Tin(IV) chloride, 3.754 g of 5-sulfoisophthalic acid (SPA) monosodium salt, and 0.476 g of imidazole (Imd) were added to 70 mL of N,N,-dimethylformamide (DMF), and stirred for 16 min. Then the solution was sealed in a 100-mL stainless-steel reactor with a Teflon liner. The reaction was carried out for 20 h at 100 °C. After being cooled to 25 °C, the synthesized PCP (SPA-Imd-TinPCP) was recovered by filtering it through a 450-nm polycarbonate membrane, and washing with DMF, distilled water, and ethanol. SPA-Imd-TinPCP was dried under vacuum at 60 °C overnight.

3.3. Characterization and Measurement

X-ray photoelectron spectroscopy (XPS) spectra were measured using an X-ray photoelectron spectrometer (Thermo VG ESCALAB250, Waltham, MA, USA). The measurement was carried out at a pressure of 2 × 10−9 Pa using Mg Ka X-ray as the excitation source. Infrared spectra were recorded on a Bruker Tensor 27 FTIR spectrometer at a nominal resolution of 2 cm−1 (Bruker, Billerica, MA, USA).
The N2 sorption measurements were performed on a Micromeritics ASAP 2020 V3.01H instrument (Micromeritics, Norcross, GA, USA) at 77 K. The samples were degassed under a high vacuum (10−5 Torr) at 150 °C for 10 h, prior to each measurement. Using the Brunauer-Emmett-Teller (BET) method, the surface area was calculated from adsorption isotherm data points and the pore size distribution was determined.
The FTIR spectra after pyridine adsorption were studied to evaluate the Lewis acid characteristics on the surface of the SPA-Imd-TinPCP. The tin cation coordinated in SPA-Imd-TinPCP was analyzed with inductively coupled plasma‒atomic emission spectroscopy (ICP-AES).

3.4. Conversion of Glucose to HMF

The glucose solution was prepared by dissolving 200 mg of glucose in 5 mL of solvent. Then 50 mg of SPA-Imd-TinPCP were added to the solution. The reaction mixture was heated to 160 °C in a 50-mL stainless-steel reactor with a Teflon liner in an oil bath under strong stirring. The reaction was stopped by cooling the reactor in an ice bath. The samples (200 μL) were centrifuged to remove all insoluble particulates and diluted with distilled water ready for analysis.

3.5. Substrate and Product Analysis

The analysis of glucose, fructose, formic acid, and levulinic acid was carried out using high-performance liquid chromatography (Shimadzu LC-10A, Kyoto, Japan) with an Aminex HPX-87H (Bio-Rad Laboratories Inc., Hercules, CA, USA) ion exclusion column.H2SO4 (5 mM) was used as the mobile phase and the flow rate was 0.6 mL/min. The column oven was 65 °C. A Diamonsil C18 column (Dikma Technologies Inc., Lake Forest, CA, USA) was used to detect HMF. A solution of methanol/water (70:30 v/v) was used as the mobile phase and the flow rate was 1.0 mL/min. The column oven temperature was 25 °C. Each reaction was carried out in triplicate and the error was analyzed.

3.6. Reuse of Catalyst

Consecutive use of SPA-Imd-TinPCP was carried out. After each run, the catalyst was recovered by filtration through a membrane, followed by washing with water and ethanol. Then the catalyst was dried at 60 °C under a vacuum overnight. The reaction time for each cycle was 5 h.

4. Conclusions

The tin-based porous coordination polymer SPA-Imd-TinPCP was synthesized using 5-sulfoisophthalic acid and imidazole as the ligands. This new material was used as a catalyst for the conversion of glucose to HMF. SPA-Imd-TinPCP possesses a multifunctional catalysis capability. The coordinated tin (IV) catalyzes the isomerization of glucose to fructose. The coordinated imidazole in SPA-Imd-TinPCP provides additional Brønsted base active sites for catalyzing glucose’s isomerization to fructose. The sulfonic group of SPA catalyzes the dehydration of fructose to HMF. Through the synergistic catalysis of the components of SPA-Imd-TinPCP, HMF yields of 59.5% in DMSO and 49.8% in H2O/THF were obtained. Consecutive use of SPA-Imd-TinPCP demonstrated that, after five reuses, there was no significant activity loss in terms of the glucose conversion and HMF yield. Aside from the catalysis advantage, SPA-Imd-TinPCP can be easily prepared at a relatively low temperature (100 °C), and calcinations are not needed, which may reduce the costs.

Author Contributions

P.J. provided the idea and design for the study; X.M. and S.S. performed the experiments; L.J. drafted the manuscript; P.J. revised it.

Funding

This research was funded by the National Science Foundation of China (21476023).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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. [Google Scholar] [CrossRef] [Green Version]
  2. Yabushita, M.; Li, P.; Islamoglu, T.; Kobayashi, H.; Fukuoka, A.; Farha, O.K.; Katz, A. Selective Metal−Organic framework catalysis of glucose to 5‑hydroxymethylfurfural using phosphate-modified NU-1000. Ind. Eng. Chem. Res. 2017, 56, 7141–7148. [Google Scholar] [CrossRef]
  3. Liu, L.; Chang, H.; Jameel, H.; Park, J.; Park, S. Catalytic conversion of biomass hydrolysate into 5‑hydroxymethylfurfural. Ind. Eng. Chem. Res. 2017, 56, 14447–14453. [Google Scholar] [CrossRef]
  4. Mamo, W.; Chebude, Y.; Márquez-Álvarez, C.; Díaz, I.; Sastre, E. Comparison of glucose conversion to 5-HMF using different modified mordenites in ionic liquid and biphasic media. Catal. Sci. Technol. 2016, 6, 2766–2774. [Google Scholar] [CrossRef]
  5. 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. [Google Scholar] [CrossRef]
  6. Yu, S.; Kim, E.; Park, S.; Song, I.K.; Jung, C. Isomerization of glucose into fructose over Mg-Al hydrotalcite catalysts. Catal. Commum. 2012, 29, 63–67. [Google Scholar] [CrossRef]
  7. Yang, Q.; Sherbahn, M.; Runge, T. Basic amino acids as green catalysts for isomerization of glucose to fructose in water. ACS Sustain. Chem. Eng. 2016, 4, 3526–3534. [Google Scholar] [CrossRef]
  8. 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. [Google Scholar] [CrossRef]
  9. Chakrabarti, A.; Sharma, M.M. Cationic ion exchange resins as catalyst. React. Polym. 1993, 20, 1–45. [Google Scholar] [CrossRef]
  10. Moliner, M.; Román-Leshkov, Y.; Davis, M.E. Tin-containing zeolites are highly active catalysts for the isomerization of glucose in water. Proc. Natl. Acad. Sci. USA 2010, 107, 6164–6168. [Google Scholar] [CrossRef] [Green Version]
  11. 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. Catal. 2014, 35, 723–732. [Google Scholar] [CrossRef]
  12. 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. [Google Scholar]
  13. 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.; et al. Metalloenzyme-like catalyzed isomerizations of sugars by Lewis acid zeolites. Proc. Natl. Acad. Sci. USA 2012, 109, 9727–9732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. 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. [Google Scholar] [CrossRef]
  15. 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. [Google Scholar] [CrossRef]
  16. 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. [Google Scholar] [CrossRef]
  17. 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. [Google Scholar] [CrossRef] [PubMed]
  18. Dijkmans, J.; Dusselier, M.; Gabriëls, D.; Houthoofd, K.; Magusin, P.C.M.M.; Huang, S.; Pontikes, Y.; Trekels, M.; Vantomme, A.; Giebeler, L.; et al. Cooperative catalysis for multistep biomass conversion with Sn/Al Beta zeolite. ACS Catal. 2015, 5, 928–940. [Google Scholar] [CrossRef]
  19. Liu, F.; Li, B.; Liu, C.; Kong, W.; Yi, X.; Zheng, A.; Qi, C. Template-free synthesis of porous carbonaceous solid acids with controllable acid sites and their excellent activity for catalyzing the synthesis of biofuels and fine chemicals. Catal. Sci. Technol. 2016, 6, 2995–3007. [Google Scholar] [CrossRef]
  20. 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. [Google Scholar] [CrossRef]
  21. 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. [Google Scholar] [CrossRef]
  22. 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. [Google Scholar] [CrossRef]
  23. 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. [Google Scholar] [CrossRef]
  24. Reche, M.T.; Osatiashtiani, A.; Durndell, L.J.; Isaacs, M.A.; Silva, Â.; Lee, A.F.; Wilson, K. Niobic acid nanoparticle catalysts for the aqueous phase transformation of glucose and fructose to 5-hydroxymethylfurfural. Catal. Sci. Technol. 2016, 6, 7334–7341. [Google Scholar] [CrossRef] [Green Version]
  25. 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. [Google Scholar] [CrossRef] [PubMed]
  26. Akiyama, G.; Matsuda, R.; Sato, H.; Takata, M.; Kitagawa, S. Cellulose hydrolysis by a new porous coordination polymer decorated with sulfonic acid functional groups. Adv. Mater. 2011, 23, 3294–3297. [Google Scholar] [CrossRef]
  27. Dong, X.Y.; Wang, R.; Wang, J.Z.; Zang, S.Q.; Mak, T.C.W. Highly selective Fe3⁺ sensing and proton conduction in a water-stable sulfonate carboxylate Tb–organic-framework. J. Mater. Chem. A 2015, 3, 641–647. [Google Scholar] [CrossRef]
  28. Ramaswamy, P.; Matsuda, R.; Kosaka, W.; Akiyama, G.; Jeon, H.J.; Kitagawa, S. Highly proton conductive nanoporous coordination polymers with sulfonic acid groups on the pore surface. Chem. Commun. 2014, 50, 1144–1146. [Google Scholar] [CrossRef]
  29. Chen, S.S. The roles of imidazole ligands in coordination supramolecular systems. CrystEngComm 2016, 18, 6543–6565. [Google Scholar] [CrossRef]
  30. Zhang, F.M.; Dong, L.Z.; Qin, J.S.; Guan, W.; Liu, J.; Li, S.L.; Lu, M.; Lan, Y.Q.; Su, Z.M.; Zhou, H.C. Effect of imidazole arrangements on proton-conductivity in metal−organic frameworks. J. Am. Chem. Soc. 2017, 139, 6183–6189. [Google Scholar] [CrossRef]
  31. Xue, G.; Dai, Q.; Jiang, S. Chemical reactions of imidazole with metallic silver studied by the use of sers and xps techniques. J. Am. Chem. Soc. 1988, 110, 2393–2395. [Google Scholar] [CrossRef]
  32. 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. [Google Scholar] [CrossRef] [PubMed]
  33. Lei, L.; Shan, J.; Hu, J.; Liu, X.; Zhao, J.; Tong, Z. Co-curing effect of imidazole grafting graphene oxide synthesized by one-pot method to reinforce epoxy nanocomposites. Compos. Sci. Technol. 2016, 128, 161–168. [Google Scholar] [CrossRef]
  34. 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. [Google Scholar] [CrossRef]
  35. 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. [Google Scholar] [CrossRef]
  36. 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. [Google Scholar] [CrossRef]
  37. Liu, W.; Wang, Q.; Zheng, Y.; Wang, S.; Yan, Y.; Yang, Y. Extraction behaviour and mechanism of Pt(IV) and Pd(II) by liquid–liquid extraction with an ionic liquid [HBBIm]Br. Dalton Trans. 2017, 46, 7210–7218. [Google Scholar] [CrossRef] [PubMed]
  38. Morzyk-Ociepa, B.; Rózycka-Sokołowska, E.; Michalska, D. Revised crystal and molecular structure, FT-IR spectra and DFT studies of chlorotetrakis (imidazole) copper(II) chloride. J. Mol. Struct. 2012, 1028, 49–56. [Google Scholar] [CrossRef]
  39. Chen, D.; Liang, F.; Feng, D.; Xian, M.; Zhang, H.; Liu, H.; Du, F. An efficient route from reproducible glucose to 5-hydroxymethylfurfural catalyzed by porous coordination polymer heterogeneous catalysts. Chem. Eng. J. 2016, 300, 177–184. [Google Scholar] [CrossRef]
  40. Kitano, M.; Wada, E.; Nakajima, K.; Hayashi, S.; Miyazaki, S.; Kobayashi, H.; Hara, M. Protonated titanate nanotubes with Lewis and Brønsted acidity: Relationship between nanotube structure and catalytic activity. Chem. Mater. 2013, 25, 385–393. [Google Scholar] [CrossRef]
  41. Atanda, L.; Mukundan, S.; Shortri, A.; Ma, Q.; Beltramini, J. Catalytic conversion of glucose to 5-hydroxymethylfurfural with a phosphated TiO2 catalyst. ChemCatChem 2015, 7, 781–790. [Google Scholar] [CrossRef]
  42. 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, 100–105. [Google Scholar] [CrossRef]
  43. 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. 2012, 14, 2506–2512. [Google Scholar] [CrossRef]
  44. Silahua-Pavóna, A.A.; Espinosa-González, C.G.; Ortiz-Chi, F.; Pacheco-Sosa, J.G.; Pérez-Vidal, H.; Arévalo-Pérez, J.C.; Godavarthi, S.; Torres-Torres, J.G. Production of 5-HMF from glucose using TiO₂–ZrO₂ catalysts: Effect of the sol-gel synthesis additive. Catal. Commun. 2019, 129, 105723. [Google Scholar] [CrossRef]
  45. Liang, F.; Chen, D.; Liu, H.; Liu, W.; Xian, M.; Feng, D. One-pot synthesis of 5‑hydroxymethylfurfural from glucose by Brønsted acid-free bifunctional porous coordination polymers in water. ACS Omega 2019, 4, 9316–9323. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Simplified reaction scheme for the conversion of glucose to HMF through glucose isomerization and subsequent dehydration.
Scheme 1. Simplified reaction scheme for the conversion of glucose to HMF through glucose isomerization and subsequent dehydration.
Catalysts 09 00739 sch001
Figure 1. TEM image (a) and XRD pattern (b) for SPA-Imd-TinPCP.
Figure 1. TEM image (a) and XRD pattern (b) for SPA-Imd-TinPCP.
Catalysts 09 00739 g001
Figure 2. (a) EDS spectrum; (b) XPS spectrum of SPA-Imd-TinPCP.
Figure 2. (a) EDS spectrum; (b) XPS spectrum of SPA-Imd-TinPCP.
Catalysts 09 00739 g002
Figure 3. XPS spectrum of N 1s of SPA-Imd-TinPCP.
Figure 3. XPS spectrum of N 1s of SPA-Imd-TinPCP.
Catalysts 09 00739 g003
Figure 4. (a) C 1s spectra for SPA-Imd-TinPCP; (b) Sn 3d spectra for SPA-Imd-TinPCP.
Figure 4. (a) C 1s spectra for SPA-Imd-TinPCP; (b) Sn 3d spectra for SPA-Imd-TinPCP.
Catalysts 09 00739 g004
Figure 5. FTIR spectrum of SPA-Imd-TinPCP.
Figure 5. FTIR spectrum of SPA-Imd-TinPCP.
Catalysts 09 00739 g005
Figure 6. The FTIR spectrum with pyridine adsorption for SPA-Imd-TinPCP.
Figure 6. The FTIR spectrum with pyridine adsorption for SPA-Imd-TinPCP.
Catalysts 09 00739 g006
Figure 7. (a) N2 adsorption and desorption isotherms measured at 77 K for SPA-Imd-TinPCP; (b) pore size distribution of SPA-Imd-TinPCP.
Figure 7. (a) N2 adsorption and desorption isotherms measured at 77 K for SPA-Imd-TinPCP; (b) pore size distribution of SPA-Imd-TinPCP.
Catalysts 09 00739 g007
Figure 8. Schematic representation of coordination involving the ligands 5-sulfoisophthanic acid and imidazole for SPA-Imd-TinPCP; tin (IV); Catalysts 09 00739 i001 imidazole; and Catalysts 09 00739 i002 5-sulfoisophthanic acid.
Figure 8. Schematic representation of coordination involving the ligands 5-sulfoisophthanic acid and imidazole for SPA-Imd-TinPCP; tin (IV); Catalysts 09 00739 i001 imidazole; and Catalysts 09 00739 i002 5-sulfoisophthanic acid.
Catalysts 09 00739 g008
Figure 9. Fructose conversion and HMF yield under the catalysis of SPA-Imd-TinPCP in DMSO (a) and H 2 O /THF (v/v 1:4) (b). The working pressures were 1 atm and 5 atm for DMSO and H2O/THF systems, respectively.
Figure 9. Fructose conversion and HMF yield under the catalysis of SPA-Imd-TinPCP in DMSO (a) and H 2 O /THF (v/v 1:4) (b). The working pressures were 1 atm and 5 atm for DMSO and H2O/THF systems, respectively.
Catalysts 09 00739 g009
Figure 10. Glucose conversion, fructose, and HMF yields versus reaction time (a) DMSO; (b) H2O/THF. The working pressures were 1 atm and 5 atm for the DMSO and H2O/THF systems, respectively.
Figure 10. Glucose conversion, fructose, and HMF yields versus reaction time (a) DMSO; (b) H2O/THF. The working pressures were 1 atm and 5 atm for the DMSO and H2O/THF systems, respectively.
Catalysts 09 00739 g010
Figure 11. Consecutive use of SPA-Imd-TinPCP for glucose conversion into HMF in (a) DMSO; (b) H2O/THF (1:4). The reaction time was 5 h. The working pressure was 5 atm.
Figure 11. Consecutive use of SPA-Imd-TinPCP for glucose conversion into HMF in (a) DMSO; (b) H2O/THF (1:4). The reaction time was 5 h. The working pressure was 5 atm.
Catalysts 09 00739 g011
Table 1. Conversion of glucose to HMF with various catalysts.
Table 1. Conversion of glucose to HMF with various catalysts.
CatalystGlucose Conversion (%)Fructose Yield (%)HMF Yield (%)
Imidazole63.2 ± 0.621.1 ± 0.31.9 ± 0.1
SPA-TinPCP70.3 ± 0.71.1 ± 0.137.3 ± 0.7
SPA-Imd-TinPCP86.2 ± 0.51.0 ± 0.149.8 ± 0.5
Table 2. Comparison of SPA-Imd-TinPCP with other catalysts: HMF yield and selectivity.
Table 2. Comparison of SPA-Imd-TinPCP with other catalysts: HMF yield and selectivity.
SolventCatalystYield (%)Selec. (%)T (°C)Time
(min)
Ref.
DMSOSnCl410.832.780180[14]
DMSOSnPO24.2 120300[22]
DMSOSn-Mont43.644.0160180[43]
DMSOSPA-Imd-TinPCP59.566.5160240this work
H2O/1-butanolSn-Beta13.518.016090[5]
H2O/THF/HClSn-Beta/HCl56.972.018070[5]
H2O/THFSPA-Imd-TinPCP49.857.8160300this work

Share and Cite

MDPI and ACS Style

Jiao, L.; Sun, S.; Meng, X.; Ji, P. Sn-Based Porous Coordination Polymer Synthesized with Two Ligands for Tandem Catalysis Producing 5-Hydroxymethylfurfural. Catalysts 2019, 9, 739. https://doi.org/10.3390/catal9090739

AMA Style

Jiao L, Sun S, Meng X, Ji P. Sn-Based Porous Coordination Polymer Synthesized with Two Ligands for Tandem Catalysis Producing 5-Hydroxymethylfurfural. Catalysts. 2019; 9(9):739. https://doi.org/10.3390/catal9090739

Chicago/Turabian Style

Jiao, Lutong, Siyu Sun, Xianling Meng, and Peijun Ji. 2019. "Sn-Based Porous Coordination Polymer Synthesized with Two Ligands for Tandem Catalysis Producing 5-Hydroxymethylfurfural" Catalysts 9, no. 9: 739. https://doi.org/10.3390/catal9090739

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop