G. Parameswaram and
Sounak Roy*
Department of Chemistry, Birla Institute of Technology and Science (BITS) Pilani, Hyderabad Campus, Jawahar Nagar, Shameerpet Mandal, Hyderabad 500078, Telangana, India. E-mail: sounak.roy@hyderabad.bits-pilani.ac.in
First published on 9th August 2018
Energy-efficient and sustainable processes for the production of 5-hydroxymethylfurfural (HMF) from carbohydrates are in high demand. Bivalent ion-exchanged microwave-synthesized ZnxTPA/γ-Al2O3 was employed for the direct conversion of carbohydrates into HMF. The as-synthesized samples were structurally characterized by FTIR and Raman spectroscopy, UV-Vis diffused reflectance spectroscopy, and X-ray diffraction. Thermal characterization was performed by TG-DTA. The surface morphology was analysed by FE-SEM, and surface area analysis was performed. The surface acidities of the as-synthesized catalysts were elucidated by pyridine FTIR spectra and NH3-TPD. The catalytic performance was thoroughly studied as a function of Zn2+ doping, reaction temperature, catalysts loading, and effect of solvents. Microwave-synthesized Zn0.5TPA/γ-Al2O3 exhibited excellent catalytic fructose dehydration, with 88% HMF yield at 120 °C for 2 h. The surface Brønsted acidity was found to be crucial for optimum catalytic activity.
Catalytic dehydration of C6 sugars to HMF is being studied by homogeneous and heterogeneous Brønsted acid catalysts. However, heterogeneous acid catalysts have become increasingly attractive due to their environmentally benign nature, in addition to their ease of separation and recycling.19–21 The most studied heterogeneous catalysts for HMF production are ion-exchange resins,22 zeolites23–25 and phosphates of transition metals (such as Nb, Zr, V, Cu, and Ti).26–30 In this context, heteropolyacids (HPAs), which are Keggin-type super acids with well-defined structures and specific Brønsted acidities, have attracted wide attention.31–36 Due to their structural characteristics and heteroatoms, it is possible to modify their acid–base properties by changing their chemical compositions, abilities to accept and release electrons, and high proton mobilities. However, HPAs generally have weak thermal stability, high solubility in polar solvents and low specific surface areas (<10 m2 g−1). Therefore, there is room for considerable improvement of HPAs towards the catalytic conversion of C6 sugars to HMF.
In this study, we have synthesized Zn2+ salts of 12-tungstophosphoric acid (TPA) supported on high surface area Al2O3 by a microwave-assisted hydrothermal method. Zinc was selected as the exchanging agent as (i) it is a Lewis acid and (ii) zinc chloride has been proved to be a useful catalyst for HMF synthesis from carbohydrates.37,38 As the introduction of Lewis acid ions in heteropolyacid molecules assists the Brønsted sites with the opportunity to design unique protons,39 the introduction of zinc to the 12-tungstophosphoric anion was expected to create a synergistic effect to obtain double acid sites with high strengths. Investigative studies exist on monovalent salts of HPA; however, there are hardly any studies on bivalent salts of HPA towards catalytic formation of HMF. To the best of our knowledge, there is also no other report on this particular novel synthesis of ZnxTPA/γ-Al2O3. The catalysts were found to provide excellent yields of HMF from fructose in dimethyl sulfoxide (DMSO) solvent. The reaction conditions were optimized, and the recyclability of the catalysts was carefully examined. The roles of the surface–structural properties of the catalysts and their influence on the dehydration of fructose have been studied in the present study.
The surface acidities were probed by FTIR spectra of pyridine adsorption over ZnxTPA/γ-Al2O3 and NH3 TPD. Pyridine was dropped on KBr pellets of ZnxTPA/γ-Al2O3, which were dried in a hot air oven; the FTIR spectra were recorded using the JASCO FTIR-4200 with a resolution of 4 cm−1 at room temperature. The NH3-TPD of the ZnxTPA/γ-Al2O3 samples was carried out in a BELCAT II instrument (Japan) in order to understand their surface acidities. About 50 mg of powdered sample was taken inside a quartz U tube and then, pre-treatment was conducted in He flow at 300 °C for 1 h prior to NH3 adsorption at 70 °C. The NH3 adsorption was continued for 45 min and then, the sample was flushed with He for 1 h at 100 °C to remove physically adsorbed NH3 from the catalyst surface. The desorption profile was recorded at a heating rate of 10 °C min−1 from 100 °C to 800 °C, and the desorbed NH3 was monitored with a thermal conductivity detector.
All reactions were monitored by liquid chromatography system (Shimadzu HPLC, Shimadzu, Japan) with a solvent delivery system of two pumps (Model LC20AD, Prominence Liquid Chromatograph, Shimadzu, Japan), an auto injector (Model SIL-20A HT, Prominence Auto Sampler, Shimadzu, Japan), and a photo diode array (PDA) UV detector (Model SPD-M20A, Prominence Diode Array Detector, Shimadzu, Japan). Data collection and integration were accomplished using LC Solutions software, version 1.25. Chromatographic separation was performed on a Thermo Scientific C18 column (dimensions: 150 mm × 4.6 mm, 5 μm) maintained at 30 °C. The mobile phase was a mixture of methanol and water (20/80 v/v). Analyses were carried out under isocratic conditions at a flow rate set at 0.8 mL min−1, and the detection wavelength was 280.0 nm. The HPLC system was stabilized for 1.5 h at a 1 mL min−1 flow rate through baseline monitoring prior to actual analysis. An injection volume of 20 μL was optimized for the final method. Fructose was quantified using an external standard at 25 °C using a Shimadzu HPLC apparatus equipped with a Shimpack GIST NH2 column (dimensions: 250 mm × 4.6 mm, 5 μm) maintained at 40 °C, a Shimadzu LC-20AD pump, and a Shimadzu RID-10A detector. The mobile phase was a mixture of acetonitrile and water (75/25 v/v). The flow rate was set at 0.8 mL min−1. The RID-10A detector was stabilized for 2 h at a flow rate of 1 mL min−1. The reaction mixture was diluted with a known volume of Milli-Q water before analysis to avoid overloading of the column. All experiments were performed three times, and the average values were reported; the standard deviations of the triplicates were <2.0%.
Fig. 1 FTIR spectra of the catalysts: (a) TPA/γ-Al2O3, (b) Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3, and (d) Zn1.5TPA/γ-Al2O3. |
The structural integrity of the Keggin unit was also investigated by Raman spectroscopy. Fig. 2 presents the Raman spectra of the TPA/γ-Al2O3 and ZnxTPA/γ-Al2O3 catalysts. PW12O403− is composed of a framework of distorted octahedral WO6 units. The characteristic Raman signal of the supported TPA observed at 985 cm−1 is ascribed to the symmetric stretching vibration of the double-bonded tungsten-terminal oxygen νs (WOt) species, along with its shoulder near 960 cm−1.40,41 The additional signal at 895 cm−1 is due to the stretching vibrations of the bridging W–O–W species of the extended polytungstate framework surrounding the central P atom. The Zn ion-exchanged TPA-supported γ-Al2O3 catalysts also exhibited all the characteristic bands associated with the Keggin ion, confirming the retention of the primary Keggin structure of heteropoly tungstate. It must be noted, however, that with increasing number of replaced protons, there was a gradual decrease in the intensity of the characteristic peaks without changing the peak position. No peak related to ZnO was observed.
Fig. 2 Laser Raman spectra of the as-prepared catalysts: (a) TPA/γ-Al2O3, (b) Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3, and (d) Zn1.5TPA/γ-Al2O3. |
The XRD patterns of the ZnxTPA/γ-Al2O3 catalysts along with TPA/γ-Al2O3 are shown in Fig. 3. The 2θ values of 10.4°, 17.93°, 23.13°, 25.42°, 29.49°, and 37.74° confirm that TPA/γ-Al2O3 crystallizes in the primitive cubic space group of Pn3m [JCPDS # 50-1857]. However, in the case of the ZnxTPA/γ-Al2O3 catalysts, the highest intensity peaks of (110) and (222) decreased drastically and additional broad peaks formed, indicating that the Zn acid salts of TPA are two-phase mixtures consisting of neutral ZnxTPA and TPA.41,42 The crystallite sizes of the catalysts were determined by applying the Debye–Scherrer formula, where D is the crystallite size, λ is the wavelength of the X-ray radiation, β is the line width and θ is the angle of diffraction to the main diffraction peaks (TPA: 2θ = 10.4°; ZnxTPA: 2θ = 9.02°). The crystallite sizes of TPA/γ-Al2O3, Zn0.5TPA/γ-Al2O3, Zn1TPA/γ-Al2O3, and Zn1.5TPA/γ-Al2O3 were found to be roughly 36, 24.6, 40.9, and 27.4 nm, respectively.
Fig. 3 XRD patterns of the catalysts: (a) TPA/γ-Al2O3, (b) Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3, and (d) Zn1.5TPA/γ-Al2O3. |
UV-Vis diffuse reflectance absorption spectra of the catalysts were obtained in order to obtain information about the chemical natures and coordination states of the microwave-assisted hydrothermally synthesized catalysts; the spectra are plotted in Fig. 4. All the as-prepared samples showed two absorption bands, one band at about 260 nm and a second band at 334 nm. The primary peak can be assigned to the oxygen metal charge transfer of tungstophosphate anion [PW12O403−].43,44 The next broad band at about 334 nm indicates the formation of some small clusters of bulk WO3, demonstrating the presence of octahedral coordination in the extra framework.45 TPA/γ-Al2O3 showed a lower intensity band compared to that of the ZnxTPA/γ-Al2O3 catalysts. The peak intensity increased with increasing Zn content. These results corroborate the data extracted from the FTIR spectroscopy, laser Raman spectroscopy, and XRD analyses, which also showed the presence of both Keggin-type polytungstate species and WO3 crystallites.
Fig. 4 Diffuse reflectance UV-Vis absorption spectra of (a) TPA/γ-Al2O3, (b)Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3, and (d) Zn1.5TPA/γ-Al2O3. |
The thermal stabilities of TPA/γ-Al2O3 and ZnxTPA/γ-Al2O3 were evaluated by TGA and DTA (Fig. 5). The graphs show typical TG/DTA patterns for Keggin-type heteropolyacids. For these materials, the initial major weight loss is related to the elimination of adsorbed water molecules. These molecules are responsible for the crystallization of heteropolyanions into a hydrate. The second weight loss may be due to the evolution of CO2 and decomposition of oxycarbonate to afford the stable catalyst. The final weight loss may be the decomposition of the heteropolyanion, with crystallization of oxides. The DTA curves show the corresponding endothermic peaks, and an exothermic peak was observed at 560 °C.
Fig. 5 DTA (left panel) and TGA (right panel) of (a) TPA/γ-Al2O3, (b) Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3, and (d) Zn1.5TPA/γ-Al2O3. |
As catalysis is primarily a surface phenomenon, the surface characteristics and surface morphologies of the as-prepared catalysts were thoroughly studied by surface area analyses, FE-SEM, and surface acidity analyses. The Langmuir surface areas, pore volumes, and pore diameters are tabulated in Table 1. With increasing Zn2+ incorporation, the surface areas and the average pore volumes gradually decreased. However, the mean pore diameter increased steadily.
Entry | Catalyst | Surface area (m2 g−1) | Pore volume (cm3 g−1) | Pore diameter (nm) |
---|---|---|---|---|
1 | Al2O3 | 173 | 0.593 | 13.76 |
2 | TPA/γ-Al2O3 | 8.68 | 0.0163 | 7.55 |
3 | Zn0.5TPA/γ-Al2O3 | 2.08 | 0.0051 | 9.88 |
4 | Zn1TPA/γ-Al2O3 | 1.6 | 0.0058 | 14.46 |
5 | Zn1.5TPA/γ-Al2O3 | 0.76 | 0.0027 | 14.36 |
The FE-SEM micrographs of the as-synthesized fresh Zn0.5TPA/γ-Al2O3 catalyst (top panel) and that exhausted after a catalytic cycle (bottom panel) are shown in Fig. 6 as an example. The freshly prepared Zn0.5TPA/γ-Al2O3 shows a small platelet-shape morphology with cracks and kinks. Also, few sub-micrometre particles were observed on top of the platelet-like morphologies. However, after one catalytic cycle, the platelets tended to agglomerate, forming a smoother surface with Zn0.5TPA/γ-Al2O3. Also, the sub-micrometre particles were no longer observed. The EDAX spectral analysis confirmed the expected Zn contents in the fresh and exhausted catalysts. The EDAX analysis of the ZnxTPA/γ-Al2O3 catalysts (x = 0.5 to 1.5) showed atomic Zn percentages of 1.39, 2.71, and 4.52 for Zn0.5TPA/γ-Al2O3, Zn1TPA/γ-Al2O3 and Zn1.5TPA/γ-Al2O3, respectively. These results are in good agreement with the ED-XRF data shown in Table 2. Apparently, Zn0.5TPA/γ-Al2O3 showed W = 95.38%, P = 1.96%, and Zn = 1.71% with a W:P:Zn atomic ratio of 12:1:0.5, corresponding to the molecular formula of Zn0.5PW12O40.
Fig. 6 FE-SEM images of Zn0.5TPA/γ-Al2O3 (the top panel represents the freshly synthesized catalyst, and the bottom panel represents the exhausted catalyst). |
S. no | Catalyst | Atomic (%) | |||
---|---|---|---|---|---|
Al | P | W | Zn | ||
1 | TPA/γ-Al2O3 | 0.40 | 1.93 | 97.65 | 0 |
2 | Zn0.5TPA/γ-Al2O3 | 0.96 | 1.96 | 95.38 | 1.71 |
3 | Zn1TPA/γ-Al2O3 | 0.95 | 1.98 | 94.21 | 2.86 |
4 | Zn1.5TPA/γ-Al2O3 | 0.95 | 1.85 | 92.57 | 4.63 |
The acidity of a catalyst is expected to play a key role in determining its catalytic performance. FTIR spectra of pyridine-adsorbed catalysts can be used to investigate the presence of Brønsted and Lewis acidic sites in super acid catalysts. The pyridine-adsorbed FTIR spectra of the TPA/γ-Al2O3 and Zn0.5TPA/γ-Al2O3 catalysts are shown in Fig. 7. It can be indicated that the main bands at 1400–1600 cm−1 are ascribed to the stretching vibrations of M–N (metal with nitrogen) and N–H (pyridinium ion). The band at 1536 cm−1 corresponds to pyridine adsorbed on Brønsted acidic sites. The other band located at 1482 cm−1 originates from a combination of pyridine bonded to both Brønsted and Lewis acidic sites.39 The next major band connected to Lewis acidic sites was clearly observed at 1441 cm−1 for the present catalysts. The intensity of the bands originating from Lewis and Brønsted acid sites increased significantly from pristine TPA/γ-Al2O3 to Zn0.5TPA/γ-Al2O3. However, with further increase in the Zn molar ratio, the bands did not increase. It can be concluded that the partial exchange of protons with Zn2+ introduces more Lewis acid sites into the catalyst and also enhances the Brønsted acid strength due to the mobility of residual protons in the secondary structure of heteropoly acid.46,47
Fig. 7 Pyridine-adsorbed FTIR spectra of the as-prepared catalysts: (a) TPA/γ-Al2O3, (b) Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3, and (d) Zn1.5TPA/γ-Al2O3. |
While the FTIR spectra of the pyridine-adsorbed catalysts show differences in their acidic sites, NH3-TPD studies semi-quantitatively determine the strengths and densities of the acid sites. The acid strengths of the pristine and Zn2+-exchanged supported TPA catalysts were studied through NH3-TPD, and the obtained results are plotted in Fig. 8. The TPD profile of pristine TPA/γ-Al2O3 shows a broad desorption peak at 100 °C to 250 °C, a small peak at 300 °C, and a sharp peak at 550 °C, indicating three different weak, moderate and strong acid sites.48 The ZnxTPA/γ-Al2O3 materials show minor differences in their TPD profiles. The weak acid sites desorbing at 100 °C to 250 °C steadily increase with the Zn2+ concentration; this may be associated with the Lewis acidic sites induced by the existence of Zn2+ in heteropolytungstate. However, the strong sites at higher temperatures initially increase up to Zn0.5 content and then decrease with increasing Zn2+. The strong acid sites may be associated with the Brønsted sites, as has been reported previously.49,50 It should be noted that Zn0.5TPA/γ-Al2O3 also shows stronger FTIR spectrum originating from Lewis and Brønsted acid sites compared to that of TPA/γ-Al2O3. A detailed observation of the TPD profiles shows that in the case of the TPA/γ-Al2O3 catalyst, there are two desorption peaks at 511 °C and 561 °C. However, with Zn2+ incorporation, the high temperature desorption peak gradually shifts to lower temperatures, indicating that with Zn contents higher than 0.5, the strong acidic sites are decreased.51 The total acidities arising from the three different sites were calculated and are plotted in Table 3. The Zn content of 0.5 shows the highest acidity value of 0.948 mmol g−1 among all the as-synthesized catalysts. The highest acidity for Zn0.5TPA/γ-Al2O3 may be due to the formation of monolayer coverage of Zn0.5TPA/γ-Al2O3. Haider et al. also observed a similar pattern of higher acidity with partially replaced protons, which they attributed to the mobility of residual protons in the secondary structure of the heteropoly acid.52
Fig. 8 NH3-TPD profiles of the as-prepared catalysts: (a) TPA/γ-Al2O3, (b) Zn0.5TPA/γ-Al2O3, (c) Zn1TPA/γ-Al2O3 and (d) Zn1.5TPA/γ-Al2O3. |
Entry | Catalyst | Acidity (mmol g−1) | Total acidity | ||
---|---|---|---|---|---|
Weak | Moderate | Strong | |||
1 | TPA/γ-Al2O3 | 0.131 | 0.025 | 0.440 | 0.596 |
2 | Zn0.5TPA/γ-Al2O3 | 0.354 | 0.0154 | 0.579 | 0.948 |
3 | Zn1TPA/γ-Al2O3 | 0.361 | 0.041 | 0.403 | 0.805 |
4 | Zn1.5TPA/γ-Al2O3 | 0.485 | 0.024 | 0.248 | 0.757 |
Entry | Catalyst | Fructose conversion (%) | HMF yield (%) |
---|---|---|---|
a Reaction conditions: fructose, 120 mg; DMSO, 2 mL; acid catalyst, 15 mg; reaction temperature, 120 °C; time, 2 h. | |||
1 | No catalyst | 81 | 2 |
2 | TPA | 76 | 70 |
3 | γ-Al2O3 | 10 | 5 |
4 | Zn0.5TPA | 94 | 73 |
5 | TPA/γ-Al2O3 | 99 | 74 |
6 | Zn0.5TPA/γ-Al2O3 | 99 | 88 |
7 | Zn1TPA/γ-Al2O3 | 99 | 83 |
8 | Zn1.5TPA/γ-Al2O3 | 99 | 25 |
The dehydration of fructose to HMF in DMSO solvent using the Zn0.5TPA/γ-Al2O3 acid catalyst can be influenced by several reaction parameters, including the reaction temperature, reaction time, catalyst weight, and nature of solvents. Therefore, further investigation was carried out to elucidate the optimum conditions of the reaction. To determine the influence of reaction temperature on the conversion of fructose to HMF, the experiments were carried out at different reaction temperatures ranging from 80 °C to 160 °C over Zn0.5TPA/γ-Al2O3 while keeping the other parameters fixed; the results are given in Fig. 9a. It can be seen that the reaction temperature had a large effect on the dehydration of fructose to HMF. At a lower temperature of 80 °C, the fructose conversion and the HMF yield were found to be the lowest. When the reaction was carried out at 100 °C, although the conversion of fructose was high, the HMF yield was as low as ∼35%. A drastic increase in the HMF yield from 35% to 88% was observed with increasing reaction temperature from 100 °C to 120 °C. However, further increasing the reaction temperature to 140 °C and 160 °C did not show any appreciable increase in the HMF yield or fructose conversion. From the economic and HMF yield points of view, 120 °C was selected as the reaction temperature of fructose dehydration reaction. Similar results were also observed in the literature.32,54 The dehydration of fructose to HMF over Zn0.5TPA/γ-Al2O3 was also studied as a function of reaction time while keeping the reaction temperature at 120 °C and the catalyst loading constant; the results are plotted in Fig. 9b. The reaction duration was varied from 30 min to 150 min. With the reaction duration of 30 min, the fructose conversion was 54% and the HMF yield was found to be only 21%. The fructose conversion increased to 77% with 37% HMF yield when the reaction was carried out for 60 min. The fructose conversion and the HMF yield increased further as the reaction time was extended to 90 min and 120 min and then remained relatively stable with further increase of the reaction time to 150 min. Thus, the optimum reaction time for the dehydration of fructose to HMF over Zn0.5TPA/γ-Al2O3 was found to be 120 min at 120 °C. The performance of the Zn0.5TPA/γ-Al2O3 catalysts was also examined with variation of the catalyst loading. Experiments with different catalyst quantities from 1 to 25 mg were conducted at 120 °C for 120 min, and the results are given in Fig. 9c. With only 1 mg of catalyst, the fructose conversion was 33%, with 11% HMF yield. With catalyst loadings of 5 and 10 mg, both the fructose conversion and product yield increased gradually. With a further increase in catalyst amount from 10 to 15 mg, 99% fructose conversion was achieved with increasing HMF yield from 35% to 88%. Further increases in catalyst loading of 20 and 25 mg resulted in slight increases in HMF yield with complete conversion of fructose. The effects of solvents other than DMSO were also studied thoroughly over our supported catalyst. For the comparison of solvents, water, ethanol and DMF were used along with DMSO. The results are plotted in Fig. 9d. Among the solvents, water showed the poorest fructose conversion and HMF yield. For ethanol and DMSO, the fructose conversion was comparable; however, ethanol showed a better HMF yield than DMF. The solvent DMSO showed the best activity. Therefore, the optimum reaction conditions for the dehydration of fructose to HMF in DMSO using Zn0.5TPA/γ-Al2O3 were found to be 120 °C for 2 h of reaction duration with a catalyst loading of 20 mg.
Fig. 10 Catalyst activity in four reaction cycles. Reaction conditions: fructose (120 mg), DMSO (2 mL), catalyst (15 mg), temperature (120 °C), time (120 min). |
Entry | Catalyst | Temperature (°C) | Time | Fructose conversion (%) | HMF yield (%) | Ref. |
---|---|---|---|---|---|---|
a Heating by microwave irradiation. | ||||||
1 | FePW12O40 | 120 | 2 h | 100 | 49 | 55 |
2 | TiO2 | 140a | 10 min | — | 54 | 57 |
3 | Amberlyst-15 | 120 | 2 h | 100 | 76 | 55 |
4 | Sn–W | 80 | 12 h | 99 | 70 | 58 |
5 | H–Y zeolite | 120 | 2 h | 100 | 76 | 55 |
6 | BEA-ZSM | 120 | 2 h | — | 78 | 59 |
7 | Nafion | 120 | 2 h | 100 | 75 | 55 |
8 | Zn0.5TPA/γ-Al2O3 | 120 | 2 h | 99 | 88 | Present work |
This journal is © The Royal Society of Chemistry 2018 |