Mesoporous ZrO2 Nanopowder Catalysts for the Synthesis of 5

Aladdin Biochem Technology Corporation. Page 4 of 30 ..... The reaction temperature and time were another critical influence factor. As shown in Figur...
0 downloads 0 Views 3MB Size
Article Cite This: ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

www.acsanm.org

Mesoporous ZrO2 Nanopowder Catalysts for the Synthesis of 5‑Hydroxymethylfurfural Yumeng Zhou,† Lijing Zhang,† and Shengyang Tao*,† †

Department of Chemistry, Dalian University of Technology, Dalian 116024, Liaoning, P.R. China

Downloaded via 178.159.97.245 on July 19, 2019 at 15:42:21 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Balancing the large surface area and high crystallinity of heterogeneous metal oxide catalysts is quite tricky. They are usually needed together to get both a large amount of active catalytic sites and high thermal stability. In this paper, the preparation of porous sulfate ZrO2 as solid acid catalysts is reported. The monolithic porous SiO2 created confined spaces in which porous ZrO2 was synthesized in it. The ZrO2 has excellent porous structure in which the specific surface area reaches 277 m2 g−1 and has good crystallinity. At the same time, they have high acid site loading (2.57 mmol g−1) upon sulfation. The sulfonic ZrO2 exhibits excellent catalytic performance when considering the dehydration of 5-HMF from D-fructose. The yield of 5-HMF is 87% for 60 min at 120 °C. Thus, this solid acid catalyst was a potential candidate for heterogeneous catalysis and biomass quick conversion. KEYWORDS: confine space, sulfated ZrO2, biomass conversion, 5-HMF



or carbon,20 have been regarded as more suitable catalysts due to their superiority of recyclability and reusability at the same time. Among them, metal oxides attract great interest because they are abundant in the earth, have a low price, and are easy to process. Alumina, zirconia, titania, and their composites show fine crystal structure, hydrothermal stability, and acidic property.21−27 Therefore, they have been regarded as ideal catalysts for applications in the biomass conversion process. However, in their natural form metal oxides exhibit low surface area and poor porosity when they have a high degree of crystallinity. Amorphous metal oxides usually have a larger surface area than crystalline ones, but their stability will be significantly reduced,28,29 which limits the application for the catalytic performance of viscous biobased molecules at high temperature. Therefore, synthesis of metal oxides with high crystallinity, large surface area, and fine porous structure is an essential job in the field of heterogeneous catalysts. On basis of the foregoing considerations, a straightforward and efficient strategy to synthesize porous zirconia (ZrO2) in monolithic porous silica is designed. As known, confined spaces can regulate the growth of the materials and their morphology. Various nanomaterials were prepared, such as nanospheres,30 spherical shells,31 and nano-polyhedron.32 In this paper, the used monolithic silica has interconnected macropores in it and mesopores in the silica skeleton. These pores act as reactors for the crystallization course of inorganic

INTRODUCTION Over the past several years, some renewable feedstocks such as biomass arouse the interest of researchers due to energy consumption. These materials are available from agricultural waste and are abundant and readily accessible.1,2 Biomass resources are considered to be the most promising alternatives to finite fossil resources, chemical intermediates, and sustainable supply of liquid fuel.3−7 5-Hydroxymethylfurfural (5-HMF) is an essential intermediate for manufacturing fine chemicals.8−10 2,5-Furan dicarboxylic acid (FDCA) can be obtained form the oxidation of 5-HMF. After hydrating 5HMF, levulinic acid (LA), ethyl levulinate (EL), and γvalerolactone (GVL) can be made into compounds. 5Hydroxymethylfurfural is hydrogenated to get 2,5-bis(hydroxymethyl)furan (BHMF), 2,5 dimethylfuran (DMF), 5-ethoxymethylfurfural (EMF), and 2,5- bis(hydroxymethyl)tetrahydrofuran (BHMTF). Therefore, the dehydration of Dfructose to 5-HMF has caused widespread concern.11 Recent years, the catalytic systems for the formation of 5-HMF from D-fructose is developing rapidly. Typically, Brønsted- and Lewis-acids are used as catalysts for this dehydration reaction. These catalysts can be used either in a homogeneous solution12,13 or as a solid acid catalyst14−16 by immobilizing the acid active site onto the surface of the support. Homogeneous acids are active catalysts with a moderate yield of about 40−60% but have severe drawbacks such as equipment corrosion and difficult product separation, which constrict a wide range of applications significantly. Heterogeneous acid catalysts, including H-form zeolites,17 heteropoly acids (HPA),18 metal oxides19 and acid-functionalized silicas © XXXX American Chemical Society

Received: May 29, 2019 Accepted: July 9, 2019 Published: July 9, 2019 A

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

Article

ACS Applied Nano Materials Scheme 1. Schematic Illustration of the Preparation of Mesoporous Sulfate Zirconia

Dehydration of D-fructose. The dehydration reaction was reacted in a 50 mL round-bottom flask with a temperature range of 100−140 °C prepared by an oil bath. The reaction was carried out at the desired temperature with 10 mL of DMSO, 100 mg of D-fructose, and 50 mg of catalyst. The samples were taken in a specific time for further HMF yield measurement. The sample was detected by highperformance liquid chromatography (HPLC) with an ultraviolet detector and differential refractive index detector, that is, Agilent poreshell 120 EC-C18 column (HPLC, Shimadzu LC-20A). Ultrapure water mixing with methanol in a 3:1 ratio was used as the eluent phase with a flow rate of 1 mL min−1. The yield of HMF was calculated according to the following equation

materials to control the size of the crystal grain. In this confined space, the ZrO2 gel transferred to crystallized ZrO2. The crystal structure and the pore morphology were ideally balanced. The following sulfated process of the zirconia endows it as abundant solid acid sites. The acid properties were studied and exploited in conjunction with tuning the acid properties. The brilliant catalytic quality and recyclability of the as-prepared sulfated zirconia in the dehydration reaction from D-fructose to 5-HMF is shown. The catalysts could be reused for more than five cycles with 80% 5-HMF yield in 120 °C.



EXPERIMENTAL SECTION

HMF Yield% =

Materials. Poly(ethylene oxide) (PEO)−PEG−PEO (F127) and polyethylene glycol (PEG, M 1/4 10 000) were obtained from SigmaAldrich Company. Absolute ethyl alcohol (analytical reagent (AR)), dimethylsulfoxide (DMSO), acetic acid (AR), concentrated sulfuric acid (AR), ammonia solution (25%), nitric acid (AR), sodium hydroxide, and concentrated hydrochloric acid (36−38%, AR) were purchased from Tianjin Kemiou Chemical Reagent Corporation. Tetramethoxysilane (TMOS) was purchased from the Chemical Factory of Wuhan University. D-fructose, 5-hydroxymethylfurfural (5HMF), zirconium(IV) n-propoxide, and commercial ZrO2 were received from the Shanghai Aladdin Biochem Technology Corporation. Synthesis of Porous ZrO2. Hierarchically porous silica (HPS) was made according to the previous report.33 PEG (8.85 g) and 30 mL of TMOS were dissolved in 75 mL of acetic acid (1M). Then the solution was stirring vigorously for 15 min in an ice bath, poured into 2 mL plastic centrifuge tubes, and aged for 36 h in a 40 °C oven. The white gel monolithic was obtained. After the treatment of aqueous ammonia in 110 °C for 9 h, the silica gel was annealed at 650 °C for 5 h. The hierarchically porous silica was obtained. F127 (3.2 g) was dissolved in 60 mL of ethanol. Four ml of hydrochloric acid and 4.6 mL of glacial acetic acid were mixed into the solution. After cooling to room temperature, zirconium(IV) n-propoxide was added to the solution and stirred vigorously for 2 h. Next the HPS monolith was dipped into the as-made ZrO2 sol for 24 h. The monolith was dried and calcined at 550 °C for 3 h. Last, the HPS template was etched twice in 90 °C 2 M NaOH. The porous ZrO2 was obtained. Preparation of Sulfonated ZrO2. The SO42−-modified ZrO2 as a representative solid acid was catalyzed for the dehydration reaction from D-fructose to 5-hydroxymethylfurfural. Hence, the functionalization of prepared porous material was carried out by the impregnation of ZrO2 in the sulfuric acid solution followed by calcination. In a typical procedure, 0.1 g of porous ZrO2 materials were immersed into 10 mL of sulfuric acid (1 mol l−1, 0.75 mol l−1, 0.5 mol l−1, and 0.25 mol l−1) and then stirred for 5 h in the room temperature. The powders were washed with DI water and dried in 100 °C. Then, the powders were annealed at 550 °C for 5 h and were named as 1.00 MZrO2/SO42−, 0.75 M-ZrO2/SO42−, 0.50 M-ZrO2/SO42−, and 0.25 MZrO2/SO42−, respectively.

moles of carbon in HMF product × 100 moles of carbon as fructose at t = 0

19

After that, the reaction solution was cooled to room temperature. Next, the catalyst was removed by a Buchner funnel and washed three times with ethanol and water followed by a dry period at 100 °C. After activation at 250 °C, the catalyst can be used for the next cycle. Characterization. Scanning electron microscope (SEM) images were obtained by a QUANTA 450 scanning electron microscope at 20 kV. The microscopic structure features of the samples were observed by transmission electron microscope (TEM) with a Tecnai F30 electron microscope which operates at 300 kV accelerating voltage and equipped with Schottky Field emission gun (FEG). The pore properties were surveyed by an ASAP 2010 analysis instrument at 77 K. The specific surface areas were calculated by the Brunauer− Emmett−Teller (BET) method, and the pore size was calculated using the Barrett−Joyner−Halenda (BJH) model. X-ray diffraction (XRD) patterns were measured on a Rigaku D/MAX-2400 X-ray powder diffraction (Japan) operating at 40 kV and 100 mA, using Cu Kα radiation. The acid features were measured by temperatureprogrammed desorption of the NH3 (NH3-TPD) with a Bjbuilder PCA-1200 and adsorbed pyridine infrared spectrum (py-IR) with a Bruker TENSOR 27 FT-IR spectrometer at a resolution of 4 cm−1 for 16 scans (1 s per scan).



RESULTS AND DISCUSSION Preparation of Porous ZrO2 Powders. The porous ZrO2 was synthesized by the template method (Scheme 1). The hierarchically porous silica (HPS) provided a confined space, and the triblock polymer F127 was used as a soft template. The coordination work of two templates formed the mesopores. The HPS was hierarchically porous. As shown in Figure 1A, the HPS monolith as a hard template had a large number of macropores. Macropores with a diameter of 1 μm were observed. The N2 adsorption−desorption isotherm of HPS showed that it had a mesoporous structure (shown in Figure S1). The specific surface area of the HPS was about 234 m2 g−1 (shown in Table 1). After impregnation, the pores of the HPS B

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

Article

ACS Applied Nano Materials

nm. The grain size was about 5 to 7 nm, according to the measurement, which was similar to the calculated results of the Scherrer Formula. XRD explored the crystal structure of porous ZrO2 materials. According to the XRD pattern (Figure 2A), ZrO2 samples calcined at different temperature were all in the tetragonal phase (JCPDS 01-080-2155) without phase transition at high temperature.34 The peaks at 30.27°, 34.81°, 50.37°, and 60.20° corresponded to the index of the crystal face of [101], [002], [200] and [211], respectively.35 With the increase of the temperature, the intensity of diffraction peaks grew stronger and the full width at half-maximum (fwhm) was reduced, meaning the crystallinity was enhanced, and the crystal grains were gradually enlarged. The crystal form did not change in HPS. According to the Scherrer Formula, the grain size grew from 4.6 to 6.6 nm as the temperature rose. Table 1 showed these results. To further investigate the porous properties of the materials, the nitrogen sorption isotherms of the samples were displyed in Figure S2. The N2 adsorption−desorption isotherm of porous ZrO2 belonged to IV-type according to IUPAC, which suggested that samples have a mesoporous structure. According to data in Table 1, the sample calcined at 450 °C had the most significant specific surface area which reached 277 m2 g−1, whereas the sample calcined at 750 °C had the lowest surface area at only 159 m2 g−1. The specific surface area had been reduced severely as calcination temperature increased. At the same time, the average pore diameter and pore volumes had decreased. The crystalline grain grew more prominent during the increasing temperature causing the specific surface areas to decrease. Unlike other samples prepared at high temperature, the sample calcined at 450 °C had a color of light gray, indicating that there were a few carbon residues in the ZrO2. Because ZrO2 prepared at 550 °C also had sizable specific surface areas, the after-treatment temperature was consequently chosen to be 550 °C. Characterization of Sulfated ZrO2 Solid Acid. Zirconia is an oxide that has both weak acidity and weak alkalinity. The sulfation process can significantly enhance the acidity of ZrO2. The ZrO2 had been immersed in a sulfuric acid solution with different concentrations and calcined to obtain the SO42− modified solid acid catalyst. As shown in Figure 2B, the functionalization of ZrO2 did not change the structure of the crystal. All the solid acids still maintained the tetragonal phase. The grain size was shown in Table 1 after calculation. The

Figure 1. SEM images of HPS (A) and the HPS filled with ZrO2 gel (B); the TEM images of the ZrO2 samples (C,D) prepared at 550 °C.

Table 1. Structural Parameters of Different ZrO2 Samples (Calcined at Different Temperature) and Different Sulfate ZrO2 Samples (Sulfated with Different Concentration Sulfuric Acid)a sample

SBET (m2 g−1)

DP (nm)

VP (cm3 g−1)

grain size (nm)

HPS 450 °C - ZrO2 550 °C - ZrO2 650 °C - ZrO2 750 °C - ZrO2 0.25M-ZrO2/SO42− 0.50M-ZrO2/SO42− 0.75M-ZrO2/SO42− 1.00M-ZrO2/SO42−

234 277 197 179 159 155 140 141 146

12.49 8.61 7.12 7.94 6.62 8.68 8.76 9.20 8.54

0.7 0.59 0.35 0.35 0.26 0.34 0.30 0.32 0.31

4.6 4.8 5.1 6.6 5.7 6.0 5.7 5.6

a

SBET is the specific surface area; DP is the diameter of pores; VP is the volume of pores; grain size is calculated by Scherrer Formula

monolith were filled with the ZrO2 gel. As shown in Figure 1C,D, TEM images confirmed that the ZrO2 samples were crystalline, and the accumulation of crystal particles caused the formation of mesopores, which had a diameter of less than 10

Figure 2. (A) The XRD patterns of different ZrO2 materials prepared at different temperature; (B) the XRD pattern of the ZrO2 and sulfate ZrO2 by different concentrations of sulfuric acid (0.25 mol L−1, 0.50 mol L−1, 0.75 mol L−1, and 1.00 mol L−1). C

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

Article

ACS Applied Nano Materials

Figure 3. (A) XPS survey spectra and (B) core excitation XPS spectra of Zr 3d;. (C) NH3-TPD curve of the ZrO2 materials after dipping in different concentrations of sulfuric acid solution. (D) The pyridine adsorption FTIR spectrum of the different sulfate zirconia.

Table 2. The Acidic Properties of the Different ZrO2/SO42−

crystal underwent secondary calcination, and the grain size increased from the previous 4.8 nm to about 5.7 nm. After the sulfate treatment, the N2 adsorption−desorption isotherms (Figure S3) all belonged to type-IV isotherm, which indicated that mesopores in the ZrO2 were not destroyed. The specific surface area of 0.25M-ZrO2/SO42−, 0.50M-ZrO2/ SO42−, 0.75M-ZrO2/SO42−, and 1.00M-ZrO2/SO42− were 155 m2 g−1, 140 m2 g−1, 141 m2 g−1, 146 m2 g−1, respectively. In comparison with the porous ZrO2 without sulfate, the decrease of surface area was perhaps due to the gathering of the grain after the secondary calcining, consistenting with the XRD results. The XPS survey and core excitation spectra of porous ZrO2 and ZrO2 with SO42− modified (0.75M-ZrO2/SO42−) are presented in Figure 3A,B. The XPS survey spectrum consisted of O 1s, Zr 3s, Zr 3p, C 1s (284 eV), and Zr 3d peaks (Figure 3A). After the sulfated treatment, an additional peak of S 2p (170 eV) was observed, which is shown in the inset of Figure 3A. The evolution of S species on the surface was explored via S 2p XPS spectra, which reveal a characteristic SO4 peak. As shown in Figure 3B, a small shift and broadening of the Zr 3d peak were detected related to the signal characteristic for sulfated zirconia (ZrO2−SO4).19 It is noted that the ZrO2 had been already modified by SO42−. The solid acid ZrO2/SO42− had been obtained. To further explore the S content of the sulfate samples, the elemental analysis was measured (Table 2). The S content of 0.25, 0.50, 0.75, and 1.00 mol L−1 was 1.44%, 1.75%, 2.02% and 1.71%, respectively. According to calculations, the sulfate content of different samples was 4.42%, 5.38%, 6.21%, and 5.23%, respectively. The acid properties were described by pyridine absorbed IR (py-IR) and temperature-programmed desorption of ammonia

sample 0.25M-ZrO2/ SO42− 0.50M-ZrO2/ SO42− 0.75M-ZrO2/ SO42− 1.00M-ZrO2/ SO42−

S content (wt %)

sulfate content (wt %)

acid loading (mmol NH3 g−1)

1.44

4.42

1.43

1.75

5.38

2.08

2.02

6.21

2.57

1.71

5.23

1.97

(NH3-TPD). TPD of ammonia is an extensively used method to characterize site densities in solid acid. As shown in Figure 3C, the spectrum showed the acid amount of the samples sulfated by different concentrations of sulfuric acid. The results are shown in Table 2. The 0.75 M sample had the most considerable acidic amount, which agreed with the element analysis results. The desorption of ammonia at 293−473 K, 473−673 K, and 673−823 K were due to weak, moderate, and strong acid sites, respectively. According to Figure 3C, the amount of weak acid peak was the most for all acid sites. The 0.75M-ZrO2/SO42− had a slight high-temperature peak for a small amount of strong acidic sites. Figure 3D showed the IR spectrum of the sample-absorbed pyridine at the region of 1400−1700 cm−1 wavenumber. The absence of a band at 1540 cm−1 is assigned to the Brønsted acidity result from surface OH groups. The bands at 1445 and 1608 cm−1, which were consistent with pyridinium ions, coordinated to Lewis acidic sites while the small band at 1540 cm−1 was ascribed to the adsorption of pyridine on Brønsted acidic sites.36 The Lewis acidic sites of 0.25, 0.50, and 0.75 mol L−1 sulfated ZrO2 samples were much higher than Brønsted acidic sites. In contrast with the porous ZrO2, Lewis acid of sulfate ZrO2 had D

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

Article

ACS Applied Nano Materials

Figure 4. (A) Effects of acid content on ZrO2 catalyzed the hydrolysis of D-fructose to 5-HMF; (D-fructose 100 mg, catalyst 50 mg, and DMSO 10 mL at 120 °C for 150 min). (B) Effects of different catalysts on the sythesis of 5-HMF; (D-fructose 100 mg, catalyst 50 mg, and DMSO 10 mL at 120 °C for 60 min). (C) Effects of reaction temperature from D-fructose to 5-HMF; (D-fructose 100 mg, DMSO 10 mL, and 0.75M-ZrO2/SO42− 50 mg). (D) Catalyst recycles in the repeat degradation reaction; (D-fructose 100 mg, DMSO 10 mL, 0.75M-ZrO2/SO42− 50 mg, and 120 °C).

Table 3. Comparison of D-Fructose Conversion Efficiency with Different Catalysts cat. 2−

0.75 M-ZrO2/SO4 SiNP-SO3H−C16 KCC-1-Pr-SO3H GIOMC-ArSO3H TfOH zeolitic Fe3O4@SiO2−SO3H

temp

time

120 120 162 140 120 100 100

1h 3h 30 min 30 min 4h 6h 2h

D-fructose/cat.

ratios

2 1.78 5 8.3 0.5 2.5 0.67

acid loading/mmol g−1

5-HMF yield

ref.

2.57 0.89

85% 87% 67.7% 91% 92% 64% 93.1%

this work 7 40 41 42 43 44

0.9 1.39

increased. The 0.75M-ZrO2/SO42− had the most Lewis acidic sites than the other samples. Catalytic Performance of ZrO2/SO42−. In previous research, the conversion of D-fructose to 5-hydroxymethylfurfural by using the sulfated zirconia solid acid catalyst is the goal. The prepared four ZrO2/SO42− catalysts as solid acids were estimated for the formation of the 5-hydroxymethylfurfural from D-fructose. To ensure optimal reaction parameters, the efficiency of temperature, time, acid content, catalyst structure, and D-fructose/catalyst ratios was explored. First, we examined the use of the acid content on ZrO2 for the dehydration of D-fructose to 5-hydroxymethylfurfural. In order to demonstrate the good catalytic properties of our synthetic materials, the performance of different catalysts were summarized in Figure 4A. The 5-HMF yield without catalyst was only 0.1%. The bare ZrO2 demonstrated low catalytic activity (150 min for 4% 5-HMF yield) for the synthesis of 5hydroxymethylfurfural due to the addition of −SO42− groups. When commercial ZrO2 (SBET = 45.78 m2 g−1, monoclinic) was used as a catalyst after sulfate treatment, the yield of the reaction was 1%. The NH3-TPD of sulfate commercial ZrO2 and porous ZrO2 without sulfate are shown in Figure S4. The

acid content of these two catalysts was 0.14 and 0.25 mmol g−1, respectively. The low yield was perhaps because of the less acidic site. As shown in Figure 4B, compared with the other sulfated zirconia catalysts, the yields of 5-HMF catalyzed by 0.75M-ZrO2/SO42− was nearly 90%, which was the highest in all catalysts. It may be due to the most significant acid content of the sample. The yield of different catalysts was 64%, 78%, 87%, 68%, respectively. The conversion of different catalysts was 66%, 80%, 88%, 71%, respectively (shown in Figure S5). The byproducts of the reaction were little. The reaction time was similar to the previous report.37 The reaction temperature and time were another critical influence factor. As shown in Figure 4C, the high temperature was beneficial for the reaction to reach the end in a short time. It is similar to the reported results.38 When the reaction performed below 110 °C, it needed more than 2 h to reach the end. The time decreased to 1 h, as the temperature increased to 120 °C. The yield of 5-HMF at 100 °C, 110 °C, 120 °C, and 130 °C were 35%, 76%, 87%, and 83%, respectively, under the catalyst of 0.75 M-ZrO2/SO42− within 1 h. Compared with other reports, 120 °C was a moderate temperature. The yield of 5-hydroxymethylfurfural increased first and then decreased E

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

ACS Applied Nano Materials



when the temperature increased at 130 °C. That should be due to the formation of byproducts.12 Meanwhile, too high of a temperature also quickly led to a byproduct reaction. Therefore, the reaction exhibits a lower yield at 130 °C. Compared with the reported efficient catalysts, the catalytic efficiency of the 0.75 M-ZrO2/SO42− catalyst was comparable and acceptable (as shown in Table 3). Compared with SiNPSO3H−C16, TfOH, and zeolitic, ZrO2/SO42− can catalyze the reaction to reach a similar or higher yield in a shorter time. Compared with KCC-1-Pr-SO3H, ZrO2/SO42− can catalyze the reaction to have a higher yield at the lower reaction temperature. The more acid loading of the ZrO2/SO42− catalyst causes the more excellent catalyst effect.39 The reusability of the catalyst is essential for reactions. As shown in Figure 4D, the catalytic performance of dehydration D-fructose to 5-HMF is still high, without a significant decrease after five times of running. The yield of 5-hydroxymethylfurfural was kept around 80%. The slight decreased reactivity could be due to the mass loss and the cover of the acid site.

CONCLUSION In summary, we presented an efficient synthetic approach for fabrication of porous zirconia. Monolithic porous silica with confined spaces in it acted as a perfect template to synthesize porous ZrO2. The crystallization course in the silica channel was controlled. The surface area and the crystallinity of ZrO2 were well balanced. The nanostructure in the ZrO2 made it easy to be sulfated and had a Lewis acid property and a high acid site loading for the dehydration reaction of D-fructose into 5-HMF. Therefore, the yield of 5-HMF reached 85% in a short time. These results proved that the method of synthesizing porous metal oxides in a confined space is a facile pathway for the development of high-performance functional materials. ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsanm.9b01008.



REFERENCES

(1) Mata, T. M.; Martins, A. A.; Caetano, N. S. Microalgae for biodiesel production and other applications: A review. Renewable Sustainable Energy Rev. 2010, 14 (1), 217−232. (2) Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41 (4), 1538−1558. (3) Huber, G. W.; Iborra, S.; Corma, A. Synthesis of transportation fuels from biomass: Chemistry, catalysts, and engineering. Chem. Rev. 2006, 106 (9), 4044−4098. (4) Nakagawa, Y.; Tamura, M.; Tomishige, K. Catalytic Reduction of Biomass-Derived Furanic Compounds with Hydrogen. ACS Catal. 2013, 3 (12), 2655−2668. (5) Climent, M. J.; Corma, A.; Iborra, S. Conversion of biomass platform molecules into fuel additives and liquid hydrocarbon fuels. Green Chem. 2014, 16 (2), 516−547. (6) Chheda, J. N.; Huber, G. W.; Dumesic, J. A. Liquid-phase catalytic processing of biomass-derived oxygenated hydrocarbons to fuels and chemicals. Angew. Chem., Int. Ed. 2007, 46 (38), 7164− 7183. (7) Yang, Z.; Qi, W.; Huang, R.; Fang, J.; Su, R.; He, Z. Functionalized silica nanoparticles for conversion of fructose to 5hydroxymethylfurfural. Chem. Eng. J. 2016, 296, 209−216. (8) Choudhary, V.; Mushrif, S. H.; Ho, C.; Anderko, A.; Nikolakis, V.; Marinkovic, N. S.; Frenkel, A. I.; Sandler, S. I.; Vlachos, D. G. Insights into the Interplay of Lewis and Brønsted Acid Catalysts in Glucose and Fructose Conversion to 5-(Hydroxymethyl)furfural and Levulinic Acid in Aqueous Media. J. Am. Chem. Soc. 2013, 135 (10), 3997−4006. (9) Wang, G.-H.; Hilgert, J.; Richter, F. H.; Wang, F.; Bongard, H.J.; Spliethoff, B.; Weidenthaler, C.; Schüth, F. Platinum-cobalt bimetallic nanoparticles in hollow carbon nanospheres for hydrogenolysis of 5-hydroxymethylfurfural. Nat. Mater. 2014, 13, 293. (10) Rosatella, A. A.; Simeonov, S. P.; Frade, R. F. M.; Afonso, C. A. M. 5-Hydroxymethylfurfural (HMF) as a building block platform: Biological properties, synthesis and synthetic applications. Green Chem. 2011, 13 (4), 754−793. (11) Tong, X.; Ma, Y.; Li, Y. Biomass into chemicals: Conversion of sugars to furan derivatives by catalytic processes. Appl. Catal., A 2010, 385 (1), 1−13. (12) Qi, L.; Mui, Y. F.; Lo, S. W.; Lui, M. Y.; Akien, G. R.; Horváth, I. T. Catalytic Conversion of Fructose, Glucose, and Sucrose to 5(Hydroxymethyl)furfural and Levulinic and Formic Acids in γValerolactone As a Green Solvent. ACS Catal. 2014, 4 (5), 1470− 1477. (13) Fachri, B. A.; Abdilla, R. M.; Bovenkamp, H. H. v. d.; Rasrendra, C. B.; Heeres, H. J. Experimental and Kinetic Modeling Studies on the Sulfuric Acid Catalyzed Conversion of D-fructose to 5Hydroxymethylfurfural and Levulinic Acid in Water. ACS Sustainable Chem. Eng. 2015, 3 (12), 3024−3034. (14) Bhanja, P.; Bhaumik, A. Porous nanomaterials as green catalyst for the conversion of biomass to bioenergy. Fuel 2016, 185, 432−441. (15) Chen, J.; Li, K.; Chen, L.; Liu, R.; Huang, X.; Ye, D. Conversion of fructose into 5-hydroxymethylfurfural catalyzed by recyclable sulfonic acid-functionalized metal-organic frameworks. Green Chem. 2014, 16 (5), 2490−2499. (16) Jiménez-Morales, I.; Moreno-Recio, M.; Santamaría-González, J.; Maireles-Torres, P.; Jiménez-López, A. Production of 5hydroxymethylfurfural from glucose using aluminium doped MCM41 silica as acid catalyst. Appl. Catal., B 2015, 164, 70−76. (17) Kruger, J. S.; Choudhary, V.; Nikolakis, V.; Vlachos, D. G. Elucidating the Roles of Zeolite H-BEA in Aqueous-Phase Fructose Dehydration and HMF Rehydration. ACS Catal. 2013, 3 (6), 1279− 1291. (18) Zhao, P.; Zhang, Y.; Wang, Y.; Cui, H.; Song, F.; Sun, X.; Zhang, L. Conversion of glucose into 5-hydroxymethylfurfural catalyzed by acid-base bifunctional heteropolyacid-based ionic hybrids. Green Chem. 2018, 20 (7), 1551−1559. (19) Osatiashtiani, A.; Lee, A. F.; Brown, D. R.; Melero, J. A.; Morales, G.; Wilson, K. Bifunctional SO4/ZrO2 catalysts for 5-





Article

Additional figures (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Shengyang Tao: 0000-0002-0567-8860 Author Contributions

The manuscript was written through the contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge the financial support from the National Natural Science Foundation of China (21872018,51703017), the Fundamental Research Funds for the Central Universities (DUT18RC(4)013), and the Financial Grant from the China Postdoctoral Science Foundation (2019T120203, 2016M601302). F

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX

Article

ACS Applied Nano Materials hydroxymethylfufural (5-HMF) production from glucose. Catal. Sci. Technol. 2014, 4 (2), 333−342. (20) Qi, X.; Guo, H.; Li, L.; Smith, R. L., Jr. Acid-Catalyzed Dehydration of Fructose into 5-Hydroxymethylfurfural by CelluloseDerived Amorphous Carbon. ChemSusChem 2012, 5 (11), 2215− 2220. (21) Khan, M. M.; Ansari, S. A.; Pradhan, D.; Ansari, M. O.; Han, D. H.; Lee, J.; Cho, M. H. Band gap engineered TiO2 nanoparticles for visible light induced photoelectrochemical and photocatalytic studies. J. Mater. Chem. A 2014, 2 (3), 637−644. (22) Liu, J.; Li, H.; Liu, Y.-C.; Lu, Y.-M.; He, J.; Liu, X.-F.; Wu, Z.B.; Yang, S. Catalytic conversion of glucose to 5-hydroxymethylfurfural over nano-sized mesoporous Al2O3-B2O3 solid acids. Catal. Commun. 2015, 62, 19−23. (23) Liu, G.; Yang, H. G.; Pan, J.; Yang, Y. Q.; Lu, G. Q.; Cheng, H. M. Titanium dioxide crystals with tailored facets. Chem. Rev. 2014, 114 (19), 9559−612. (24) Dutta, B.; Biswas, S.; Sharma, V.; Savage, N. O.; Alpay, S. P.; Suib, S. L. Mesoporous Manganese Oxide Catalyzed Aerobic Oxidative Coupling of Anilines To Aromatic Azo Compounds. Angew. Chem., Int. Ed. 2016, 55 (6), 2171−2175. (25) Rosen, J.; Hutchings, G. S.; Jiao, F. Ordered mesoporous cobalt oxide as highly efficient oxygen evolution catalyst. J. Am. Chem. Soc. 2013, 135 (11), 4516−4521. (26) Sayle, T. X.; Sayle, D. C. Visualizing the enhanced chemical reactivity of mesoporous ceria; simulating templated crystallization in silica scaffolds at the atomic level. J. Am. Chem. Soc. 2014, 136 (10), 4056−4065. (27) Platero-Prats, A. E.; Mavrandonakis, A.; Gallington, L. C.; Liu, Y.; Hupp, J. T.; Farha, O. K.; Cramer, C. J.; Chapman, K. W. Structural Transitions of the Metal-Oxide Nodes within MetalOrganic Frameworks: On the Local Structures of NU-1000 and UiO66. J. Am. Chem. Soc. 2016, 138 (12), 4178−4185. (28) Parlett, C. M. A.; Wilson, K.; Lee, A. F. Hierarchical porous materials: catalytic applications. Chem. Soc. Rev. 2013, 42 (9), 3876− 3893. (29) Rudisill, S. G.; Wang, Z.; Stein, A. Maintaining the Structure of Templated Porous Materials for Reactive and High-Temperature Applications. Langmuir 2012, 28 (19), 7310−7324. (30) Zhou, J.; Qin, J.; Zhang, X.; Shi, C.; Liu, E.; Li, J.; Zhao, N.; He, C. 2D Space-Confined Synthesis of Few-Layer MoS2 Anchored on Carbon Nanosheet for Lithium-Ion Battery Anode. ACS Nano 2015, 9 (4), 3837−3848. (31) Li, Q.; Wang, F.; Sun, L.; Jiang, Z.; Ye, T.; Chen, M.; Bai, Q.; Wang, C.; Han, X. Design and Synthesis of Cu@CuS Yolk-Shell Structures with Enhanced Photocatalytic Activity. Nano-Micro Lett. 2017, 9 (3), 1−9. (32) Jing, X.; He, C.; Yang, Y.; Duan, C. A Metal-Organic Tetrahedron as a Redox Vehicle to Encapsulate Organic Dyes for Photocatalytic Proton Reduction. J. Am. Chem. Soc. 2015, 137 (11), 3967−3974. (33) Wang, Y.; Shi, D.; Tao, S.; Song, W.; Wang, H.; Wang, X.; Li, G.; Qiu, J.; Ji, M. A General, Green Chemistry Approach for Immobilization of Inorganic Catalysts in Monolithic Porous FlowReactors. ACS Sustainable Chem. Eng. 2016, 4 (3), 1602−1610. (34) Wang, Y.; Liu, Y.; Yang, L.; Ruan, R.; Wen, P.; Wan, Y. Syntheses of 5-Hydroxymethylfurfural Through Glucose Dehydration in Diphasic Solvent System on ZrO2 and SO42‑/TiO2-SiO2 Catalyst. Synth. React. Inorg., Met.-Org., Nano-Met. Chem. 2016, 46 (2), 177− 184. (35) Majedi, A.; Davar, F.; Abbasi, A. Sucrose-mediated sol-gel synthesis of nanosized pure and S-doped zirconia and its catalytic activity for the synthesis of acetyl salicylic acid. J. Ind. Eng. Chem. 2014, 20 (6), 4215−4223. (36) He, J.; Chen, J.; Ren, L.; Wang, Y.; Teng, C.; Hong, M.; Zhao, J.; Jiang, B. Fabrication of Monodisperse Porous Zirconia Microspheres and Their Phosphorylation for Friedel-Crafts Alkylation of Indoles. ACS Appl. Mater. Interfaces 2014, 6 (4), 2718−2725.

(37) Ngee, E. L. S.; Gao, Y.; Chen, X.; Lee, T. M.; Hu, Z.; Zhao, D.; Yan, N. Sulfated Mesoporous Niobium Oxide Catalyzed 5Hydroxymethylfurfural Formation from Sugars. Ind. Eng. Chem. Res. 2014, 53 (37), 14225−14233. (38) Xu, H.; Miao, Z.; Zhao, H.; Yang, J.; Zhao, J.; Song, H.; Liang, N.; Chou, L. Dehydration of fructose into 5-hydroxymethylfurfural by high stable ordered mesoporous zirconium phosphate. Fuel 2015, 145, 234−240. (39) Ramli, N. A. S.; Amin, N. A. S. A new functionalized ionic liquid for efficient glucose conversion to 5-hydroxymethyl furfural and levulinic acid. J. Mol. Catal. A: Chem. 2015, 407, 113−121. (40) Najafi Chermahini, A.; Shahangi, F.; Dabbagh, H. A.; Saraji, M. Production of 5-hydroxymethylfurfural from fructose using a spherically fibrous KCC-1 silica catalyst. RSC Adv. 2016, 6 (40), 33804−33810. (41) Karimi, B.; Mirzaei, H. M.; Behzadnia, H.; Vali, H. Novel Ordered Mesoporous Carbon Based Sulfonic Acid as an Efficient Catalyst in the Selective Dehydration of Fructose into 5-HMF: the Role of Solvent and Surface Chemistry. ACS Appl. Mater. Interfaces 2015, 7 (34), 19050−9. (42) 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. (43) Mamo, W.; Chebude, Y.; Márquez-Á lvarez, 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 (8), 2766−2774. (44) Wang, S.; Zhang, Z.; Liu, B. Catalytic Conversion of Fructose and 5-Hydroxymethylfurfural into 2,5-Furandicarboxylic Acid over a Recyclable Fe3O4-CoOxMagnetite Nanocatalyst. ACS Sustainable Chem. Eng. 2015, 3 (3), 406−412.

G

DOI: 10.1021/acsanm.9b01008 ACS Appl. Nano Mater. XXXX, XXX, XXX−XXX