Catalytic Advances in the Production and Application of Biomass

Feb 28, 2018 - Chemical Transformations of Biomass-Derived C6-Furanic Platform Chemicals for Sustainable Energy Research, Materials Science, and ...
0 downloads 0 Views 5MB Size
Review Cite This: ACS Catal. 2018, 8, 2959−2980

pubs.acs.org/acscatalysis

Catalytic Advances in the Production and Application of BiomassDerived 2,5-Dihydroxymethylfuran Lei Hu,*,† Jiaxing Xu,† Shouyong Zhou,† Aiyong He,† Xing Tang,‡ Lu Lin,‡ Jiming Xu,† and Yijiang Zhao† †

Jiangsu Key Laboratory for Biomass-Based Energy and Enzyme Technology, Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection, School of Chemistry and Chemical Engineering, Huaiyin Normal University, Huaian 223300, China ‡ College of Energy, Xiamen University, Xiamen 361102, China ABSTRACT: In recent years, 2,5-dihydroxymethylfuran (DHMF), which can be produced by the selective hydrogenation of biomass-derived 5-hydroxymethylfurfural (HMF), has attracted great attention and interest of many scientists because of its peculiar symmetrical structure and wide potential applications. At present, studies of the production of DHMF are quickly progressing, with productive approaches being increasingly developed, and many crucial achievements have been continually obtained. However, to date, a special and real-time review of this research area is still lacking. To gain more insight into the current research situation, this review comprehensively summarizes and discusses state-of-theart advancements of the production of DHMF from HMF via various chemocatalytic pathways, such as conventional hydrogenation, transfer hydrogenation, electrocatalytic hydrogenation, photocatalytic hydrogenation, disproportionation reaction, and biocatalytic pathways. Meanwhile, this review also systematically outlines the latest results on the further transformation of DHMF into value-added derivatives via etherification, polymerization, and rearrangement. KEYWORDS: biomass, 5-hydroxymethylfurfural, 2,5-dihydroxymethylfuran, selective hydrogenation, further transformation, value-added derivatives hexanediol (HDO),103−107 and 3-hydroxymethylcyclopentanone (HMCPN).108−110 Among them, DHMF is highly attractive because of its peculiar symmetrical structure and wide potential applications in the production of ethers,111−113 ketones,114 and polymers;115−120 hence, it has received much more attention and interest from various scientists. It is generally known that HMF contains CO, CC, and C−O bonds that cause its very strong reactivity.121−125 Thus, ensuring the hydrogenation of the CO bond and avoiding the further hydrogenolysis of CC and C−O bonds are the key issues in the selective hydrogenation of HMF into DHMF (Figure 2), which is very similar to the exclusive reduction of α,β-unsaturated aldehydes or ketones into the corresponding unsaturated alcohols.126−131 To solve these issues, many chemocatalytic hydrogenation pathways, such as conventional hydrogenation, transfer hydrogenation, electrocatalytic hydrogenation, photocatalytic hydrogenation, disproportionation reaction, and biocatalytic hydrogenation pathways, have been gradually developed in recent years. In 2013 and 2016, Nakagawa et al.121 and Gilkey and Xu132 reviewed some of

1. INTRODUCTION At present, most chemicals and fuels in the world are directly or indirectly produced from nonrenewable fossil resources.1−5 However, as the depletion of fossil resources continues, exploring appropriate renewable resources has become increasingly urgent and necessary.6−10 As the only carbonbased renewable resource in the nature, biomass, possessing many excellent merits such as abundance, diversity, pervasiveness, and low cost, is considered to be an optimal and inexhaustible feedstock for the sustainable production of chemicals and fuels.11−16 To make good use of biomass, biorefining is a very momentous approach.17−20 During the biorefining process, 5-hydroxymethylfurfural (HMF), which is obtained by the dehydration of biomass-derived carbohydrates such as fructose,21−26 glucose,27−32 sucrose,33−38 cellobiose,39−43 inulin,44−48 starch,49−54 and cellulose,55−60 is hailed as one of the most important fundamental compounds61−65 because it can be used to synthesize many high-value products via various reactions (Figure 1). For instance, the selective hydrogenation of HMF can generate 2,5-dihydroxymethylfuran (DHMF), 6 6 − 7 0 2,5-dihydroxymethyltetrahydrofuran (DHMTHF),71−76 2,5-dimethylfuran (DMF),77−88 2,5-dimethyltetrahydrofuran (DMTHF),89−92 1-hydroxyhexane-2,5dione (HHD),93−99 1,2,6-hexanetriol (HTO),100−102 1,6© 2018 American Chemical Society

Received: October 16, 2017 Revised: February 23, 2018 Published: February 28, 2018 2959

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

Figure 1. Catalytic conversion of biomass-derived HMF into various valuable products. DFF, FDCA, FFCA, MA, FDMC, HHMFO, DMF, DMTHF, DHMTHF, HHD, HTO, BHMF, AAMFM, AOOMF, and OBMF denote 2,5-diformylfuran, 2,5-furandicarboxylic acid, 5-formyl-2furancarboxylic acid, maleic anhydride, 2,5-furandimethylcarboxylate, 5-hydroxy-5-(hydroxymethyl)furan-2(5H)-one, 2,5-dimethylfuran, 2,5dimethyltetrahydrofuran, 2,5-dihydroxymethyltetrahydrofuran, 1-hydroxyhexane-2,5-dione, 1,2,6-hexanetriol, 5,5-bis(hydroxymethyl)furoin, 5arylaminomethyl-2-furanmethanol, 5-alkanoyloxymethylfurfural, and 5,5-oxybis(methylene-2-furaldehyde), respectively.

2. CHEMOCATALYTIC HYDROGENATION OF HMF INTO DHMF In earlier studies, DHMF was mainly prepared by the stoichiometric hydrogenation of HMF in the presence of lithium aluminum hydride (LiAlH4) or sodium borohydride (NaBH4).133−137 However, this pathway has some problems such as the post-treatment of reducing agents and the coproduction of equivalent salts. To overcome these problems, substantial efforts have been made to find more proper and greener chemocatalytic pathways that are able to effectively and selectively hydrogenate HMF into DHMF. In the subsequent sections, we summarize various chemocatalytic pathways, including conventional hydrogenation, transfer hydrogenation, electrocatalytic hydrogenation, photocatalytic hydrogenation, and disproportionation reaction, in a comprehensive manner, and we also discuss the effects of various reaction parameters, including hydrogen donors, catalysts, and solvents, on the selective hydrogenation of HMF into DHMF. 2.1. Conventional Hydrogenation Pathway. As is known to all, employing molecular hydrogen (H2) as a

the available catalytic pathways for the transformation of biomass-derived fundamental compounds, such as glycerol (GC), furfural (FF), and levulinic acid (LA), in which the selective hydrogenation of HMF into DHMF is partially involved. However, to the best of our knowledge, a special and real-time review of this research area has not been published until now. To gain more insight into the current status of research, state-of-the-art advancements of the catalytic pathways for the selective hydrogenation of HMF into DHMF are comprehensively summarized and discussed in this review. Furthermore, the latest results on the further transformation of DHMF into other valuable derivatives via etherification, polymerization, and rearrangement are also systematically outlined. In a word, the main aim of this review is to attract attention to DHMF and provide some theoretical references and technical support for the practical production and application of DHMF in the near future. 2960

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

catalyst displayed good catalytic performance for the selective hydrogenation of HMF into DHMF, it is a homogeneous catalyst and shows poor recyclability. In this context, heterogeneous catalysts have drawn wide attention (Table 1). Recently, a series of ruthenium-based heterogeneous catalysts, such as carbon-supported ruthenium (Ru/C),139 zirconia-supported ruthenium (Ru/ZrO2),140 alumina-supported ruthenium (Ru/Al2O3),141 cerium oxide-supported ruthenium (Ru/CeOx),141 zirconia−magnesia-supported ruthenium (Ru/ZrO2−MgO),141 and zirconia−silica-supported ruthenium (Ru/ZrO2−SiO2),142 were designed and prepared. When they were applied for the selective hydrogenation of HMF in water, 1-butanol, or 1-butanol/water, DHMF was generated in high yields of 74.5−100% under moderate reaction conditions. More importantly, these ruthenium-based heterogeneous catalysts could be readily isolated from the corresponding catalytic systems by filtration and then successively used at least five times without an obvious loss in the yield of DHMF,139−142 clearly indicating that they also possessed superior retrievability and reusability. Unsurprisingly, like ruthenium-based heterogeneous catalysts, platinum-based heterogeneous catalysts exhibited parallel catalytic abilities for the selective hydrogenation of HMF. When carbon-supported platinum (Pt/C), 143 alumina-supported platinum (Pt/ Al2O3),112 and zeolite-supported platinum (Pt/MCM-41)144 were used as catalysts, DHMF could be obtained in 82−100% yield in water or ethanol. Inspired by these satisfactory results, other heterogeneous catalysts such as carbon-supported palladium (Pd/C),145 titania-supported iridium (Ir/TiO2),146 and iron oxide-supported gold (Au/FeOx)147 were subsequently synthesized and employed for the selective hydrogenation of HMF into DHMF. Notably, most of these heterogeneous catalysts are composed of active metals and auxiliary supports. Apart from the metal properties, such as electron deficiency, particle size, and degree of dispersion, the selective hydrogenation of HMF into DHMF may be strongly affected by the nature of the support. To verify this speculation, Ohyama et al.148 prepared a variety of gold catalysts. The experimental results demonstrated that when gold was immobilized on basic supports, such as alumina (Al2O3), lanthana (La2O3), and ceria (CeO2), the corresponding catalysts could effectively catalyze the selective hydrogenation of HMF into DHMF. However, when gold was immobilized on acidic supports, such as titania (TiO2), tantalum pentoxide (Ta2O5), and sulfated zirconia (SO42−/ZrO2), almost no

Figure 2. Selective hydrogenation of HMF into DHMF. MF, MFA, and HDO denote 5-methylfurfural, 5-methylfurfuryl alcohol, and 1,6hexanediol, respectively.

hydrogen donor for the selective hydrogenation of HMF is the most common and important pathway to produce DHMF, and it has been widely studied over various catalysts in recent years. For example, a ruthenium-based dinuclear complex (Ru-based Shvo’s catalyst) was used by Pasini et al.138 in 2014 for the selective hydrogenation of HMF. Under 10 bar H2, an excellent 99% yield of DHMF was achieved in the presence of toluene at 90 °C for 1 h. Interestingly, the combined experimental and theoretical results indicated that the hydroxyl group moiety in HMF plays an important role in the catalytic cycle toward the formation of various intermediates from HMF to DHMF (Figure 3). More surprisingly, a very simple method for separating the target product was developed by directly cooling the reactor to room temperature, which led to the quantitative precipitation of DHMF from the reaction mixture.138 In addition, it should be noted that although the Ru-based Shvo’s

Figure 3. Possible intermediates in the selective hydrogenation of HMF into DHMF over Ru-based Shvo’s catalyst. Adapted with permission from ref 138. Copyright 2014 Royal Society of Chemistry. 2961

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis Table 1. Conventional Hydrogenation of HMF into DHMF Using H2 as a Hydrogen Donor catalyst

solvent

pressure (bar)

temperature (°C)

time (h)

HMF conversion (%)

DHMF yield (%)

ref

Shvo’s Ru/C Ru/ZrO2 Ru/Al2O3 Ru/CeOx Ru/ZrO2−MgO Ru/ZrO2−SiO2 Pt/C Pt/Al2O3 Pt/MCM-41 Pd/C Ir/TiO2 Au/FeOx Au/Al2O3 Cu/C Cu/SiO2 Cu/PMO Cu/ZnO Raney Cu Raney Ni NiFe/CNT CuNi/Al2O3 CuZn/C CuZn CoAl PtCo/HCS IrRe/SiO2 PtSn/SnO2/RGO

toluene water 1-butanol 1-butanol/water 1-butanol/water 1-butanol/water water ethanol ethanol water tetrahydrofuran/water water water water ethanol methanol ethanol 1,4-dioxane water water 1-butanol tetrahydrofuran ethanol ethanol methanol 1-butanol water ethanol

10 50 15 27 27 27 5 14 14 8 100 60 30 65 50 25 50 15 90 90 30 30 50 70 40 10 8 20

90 60 120 130 130 130 25 23 23 35 80 50 80 120 180 100 100 100 90 90 110 130 180 120 120 120 30 70

1 2/3 6 2 2 2 4 18 18 2 20 3 2 2 8 8 3 2 8 8 18 6 8 3 4 2 6 1/2

99 100 99 92 100 99 98.1 − − 100 97 99 96 100 70 100 100 100 94 100 100 70.6 63.9 100 89.4 100 99 99

99 100 99 74.5 81 93.1 90.4 82 85 98.9 82 94.9 96 96 53.8 97 99 99.1 86.5 60 96.1 62.4 52.2 95 83 70 99 99

138 139 140 141 141 141 142 112 112 144 145 146 147 148 152 113 149 150 151 151 153 154 152 155 156 157 158 159

be applied for the synthesis of other products such as DMF and DHMTHF.149−152 In contrast to monometallic heterogeneous catalysts, bimetallic heterogeneous catalysts are thought to be more effective for the selective hydrogenation of HMF into DHMF because of their special geometric structures and electronic environments.152−156 For instance, carbon-nanotube-supported nickel and iron (NiFe/CNT), developed by Yu et al.,153 provided a 96.1% yield of DHMF at 110 °C for 18 h in the presence of 30 bar H2, whereas the DHMF yield was only 76.4% over carbon-nanotube-supported nickel (Ni/CNT). More interestingly, when carbon-nanotube-supported iron (Fe/CNT) was used, DHMF was not produced under the same reaction conditions, evidently demonstrating that the synergy between Ni and Fe was the main reason for the excellent catalytic performance of NiFe/CNT, which could be proved by the results of density functional theory (DFT) calculations. To be specific, the addition of oxyphilic Fe on the basis of Ni could form a NiFe alloy that was very beneficial to the stabilization of the surface species of η2-(C,O) coordination, which would further promote the selective hydrogenation of CO of HMF on the bimetallic alloy as a result of the strong interaction between the carbonyl O and the oxyphilic Fe.153 Motivated by NiFe/CNT, alumina-supported copper and nickel (CuNi/Al2O3),154 carbon-supported copper and zinc (CuZn/C),152 copper−zinc nanoalloy (CuZn),155 and cobalt− aluminum nanoalloy (CoAl)156 were also prepared for the selective hydrogenation of HMF; comparable yields of DHMF were achieved under similar reaction conditions to those over NiFe/CNT. Moreover, the bimetallic heterogeneous catalysts containing precious metals could allow milder reaction

DHMF was formed under the same reaction conditions, suggesting that compared with acidic supports, basic supports are more conducive to the formation of DHMF, probably because they favor the adsorption of CO and the dissociation of H2.148 Hence, it can be concluded that the remarkable catalytic performance of heterogeneous catalysts for the selective hydrogenation of HMF into DHMF should be ascribed to the synergistic actions of the active metals and auxiliary supports. In the above heterogeneous catalysts, Ru, Pt, Pd, Ir, and Au are precious metals and therefore are very expensive. From the viewpoint of economization and industrialization, the development of nonprecious-metal-based heterogeneous catalysts for the selective hydrogenation of HMF into DHMF is highly desirable. To explore the potential of nonprecious-metal-based heterogeneous catalysts, Cao et al.,113 Kumalaputri et al.,149 Zhu et al.,150 and Lima et al.151 prepared silica-supported copper (Cu/SiO2), copper-doped porous metal oxide (Cu/ PMO), zinc oxide-supported copper (Cu/ZnO), Raney Cu, and Raney Ni, respectively. When they were employed as catalysts for the selective hydrogenation of HMF in methanol, ethanol, 1,4-dioxane, and water, the yields of DHMF were 97, 99, 99.1, 86.5, and 60% under 25, 50, 15, 90, and 90 bar H2 at 100, 100, 100, 90, and 90 °C for 8, 3, 2, 8, and 8 h, respectively, indicating that the nonprecious-metal-based heterogeneous catalysts, especially the Cu-based heterogeneous catalysts, possessed catalytic performance similar to that of preciousmetal-based heterogeneous catalysts for the selective hydrogenation of HMF into DHMF. More gratifyingly, by adjusting the hydrogen pressure and reaction temperature, these nonprecious-metal-based heterogeneous catalysts could also 2962

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

value of the solvent has a great impact on the conversion of the substrate.144 Compared with polar solvents, nonpolar solvents exhibited worse performance for the selective hydrogenation of HMF into DHMF under the same conditions, which might be caused by a lower solubility of H2 and more difficult polarization of CO in nonpolar solvents.140 Additionally, the pH of the solution is also a momentous factor influencing the selective hydrogenation of HMF into DHMF. In contrast to a solution at pH 7, when this reaction was conducted in a solution with pH < 7, HMF could be nearly completely converted, but the selectivity for DHMF was unsatisfactory because of the formation of undesirable byproducts, such as 1,2,5-hexanetriol (1,2,5-HT), 1,2,6-hexanetriol (1,2,6-HT), and 1,2,5,6-hexanetetrol (1,2,5,6-HT), resulting from the acidcatalyzed ring opening and subsequent hydrogenation of DHMF.141 When this reaction was conducted in a solution with pH > 7, DHMF was generated with high selectivity; however, the conversion of HMF was very low.144 Accordingly, it is concluded from the above analysis that solvents with smaller δ values and stronger polarity at neutral pH are regarded as good solvents, and such solvents have been widely used by many researchers for the selective hydrogenation of HMF into DHMF (Table 1). By the conventional catalytic pathway, HMF can be selectively hydrogenated into DHMF, in which H2 is mostly employed as the hydrogen donor because of its wide availability and easy activation on many metal surfaces. However, the low solubility of H2 in most solvents requires a high hydrogen pressure to achieve the desired conversion and yield. Apart from the safety issue, handling high-pressure hydrogen will incur high infrastructure costs, which introduces an economic barrier to the production of DHMF. To further improve the security and economy of the selective hydrogenation of HMF into DHMF, many novel catalytic pathways without the addition of external molecular H2 have been gradually developed in recent years. 2.2. Transfer Hydrogenation Pathway. The pathway of catalytic transfer hydrogenation (CTH) offers a very promising alternative to the conventional hydrogenation pathway with external molecular H2. In particular, carbonyl compounds,

conditions for the selective hydrogenation of HMF into DHMF.157−159 In particular, when silica-supported iridium and rhenium (IrRe/SiO2)158 and graphene oxide-supported platinum and stannum (PtSn/SnO2/RGO)159 were used for the selective hydrogenation of HMF, DHMF was effectively synthesized in high yields of 99 and 99% at very low temperatures of 30 and 70 °C in 6 and 0.5 h under 8 and 20 bar H2, respectively. In addition to the type of catalyst, the selective hydrogenation of HMF into DHMF is also highly associated with the properties of the solvent. Generally, the solvents employed in the selective hydrogenation of HMF into DHMF can be divided into polar protic solvents (e.g., water, methanol, ethanol, propanol, and butanol), polar aprotic solvents (e.g., acetone and tetrahydrofuran), and nonpolar solvents (e.g., hexane), and they have various δ values representing the difference between the acceptor and donor numbers, which are related to quantitative measures of the Lewis acidity and basicity, respectively. Because of their negative δ values, the protic solvents, which are capable of accepting electrons, showed better performance than the aprotic solvents with positive δ values that are the symbol of donating electrons, as could be verified by the results of Chatterjee et al.144 and Han et al.140 Furthermore, with an increase in the value of δ, a declining tendency in the conversion of HMF was also observed in protic solvents (Figure 4), indicating that the δ

Figure 4. Relationship between the conversion of HMF and the δ value of the solvent over Pt/MCM-41. Adapted with permission from ref 144. Copyright 2014 Royal Society of Chemistry.

Table 2. Transfer Hydrogenation of HMF into DHMF Using Formic Acid and Alcohols as Hydrogen Donors hydrogen donor

catalyst

solvent

temperature (°C)

time (h)

HMF conversion (%)

DHMF yield (%)

ref

formic acid formic acid formic acid formic acid formic acid ethanol isopropanol isopropanol isopropanol isopropanol isopropanol isopropanol isopropanol isopropanol 2-butanol methanol 1,4-butanediol benzyl alcohol

Cm*Ru(HTsDPEN) Cp*Ir(HTsDPEN) Cp*Ir(HTsDACH) Cp*Ir(NHCPh2C6H4) Pd/C ZrO(OH)2 Zr-ATPN Zr-BPPN Zr-FDCN Fe2O3@HAP Co3O4@MC Au/SiC Ru/Co3O4 Ru/ZnAlZr-LDH MZH(Zr/Fe = 2) MgO Cu/AlOx RuCo/C

tetrahydrofuran tetrahydrofuran tetrahydrofuran tetrahydrofuran tetrahydrofuran ethanol isopropanol isopropanol isopropanol isopropanol isopropanol isopropanol isopropanol isopropanol 2-butanol methanol 1,4-butanediol benzyl alcohol

40 40 40 40 70 150 140 120 140 180 140 20 190 200 150 160 220 150

2 2 1 1 4 5/2 2 2 8 10 12 4 6 1/2 5 3 1/100 10

100 100 100 100 − 94.1 99 99 100 78.2 100 90 100 100 98.4 100 94 90.7

99 99 99 99 94 83.7 98 93 87 72 97 83.7 82.8 94 89.6 100 93 86.9

199 199 199 199 200 203 204 205 206 207 208 209 210 163 215 216 137 171

2963

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

heterogeneous catalyst for the selective hydrogenation of HMF, a comparable DHMF yield of 94% was obtained under similar reaction conditions.200 More importantly, Pd/C had good recyclability, which could further increase the workability of CTH in the presence of formic acid. However, formic acid has strong acidity and corrosivity, and its adoption requires acid- and corrosion-resistant equipment. Hence, taking this issue into account, the practical application of formic acid as a hydrogen donor may be restricted in the selective hydrogenation of HMF into DHMF. To avoid the deficiencies of formic acid, ethanol and isopropanol were subsequently investigated as hydrogen donors.201−211 In the selective hydrogenation of HMF into DHMF, alcohol-based CTH is generally slower than the H2based hydrogenation process, but it has still been found to be an effective pathway that can proceed through direct hydrogen transfer rather than the metal hydride route.132 To gain more insight into this process, we prepared a low-cost zirconium hydroxide (ZrO(OH)2) for the selective hydrogenation of HMF in the presence of ethanol.203 After 2.5 h at 150 °C, an 83.7% yield of DHMF was obtained. By combining the results of catalyst characterizations and poisoning experiments, the zirconium centers and hydroxyl groups of ZrO(OH)2 as the acidic and basic sites, respectively, were proved to play a synergistic role in the direct hydrogen transfer (Figure 6).

including aldehydes and ketones, can be exclusively reduced to the corresponding alcohols via the Meerwein−Ponndorf− Verley (MPV) reaction by using formic acid and alcohols as hydrogen donors.160−164 Compared with H2, formic acid and alcohols such as ethanol and isopropanol usually exist in the liquid form under ambient conditions, and therefore, they are more convenient and secure in the processes of storage, transportation, and usage. Furthermore, when they are used as hydrogen donors, additional facilities for the delivery of H2 are unnecessary, leading to a decrease in the number of unit operations. In addition, alcohols can also act as reaction media. Hence, when alcohols are used as hydrogen donors, additional reaction media are no longer needed, which can also enhance the economy of CTH to some extent. If the alcohol is cheap enough, the economy of CTH will be more obvious. More importantly, after the end of CTH, the unconverted alcohol can be separated from the reaction mixture and then reused as a hydrogen donor and reaction medium, and the converted alcohol, that is, the produced aldehyde or ketone, can also be separated and then integrated into the other reaction steps, such as the carbon chain growth reaction via aldol condensation. On the basis of the above-described advantages, the pathway of CTH has long been extensively employed for the selective reduction of various carbonyl compounds, such as citral (CT),165−168 FF,169−173 benzaldehyde (BA),174−178 crotonaldehyde (CA), 179−183 LA, 184−188 and ethyl levulinate (EL),189−193 via the MPV reaction. In this respect, it is necessary to emphasize that HMF is also a carbonyl compound with an aldehyde group, and it should be applicable to CTH in theory (Table 2). To verify the feasibility and availability of CTH in the selective hydrogenation of HMF, formic acid, which is an important and renewable coproduct in the acidcatalyzed decomposition of biomass-derived carbohydrates into LA,194−198 was first adopted as a hydrogen donor by Thananatthanachon and Rauchfuss.199 Surprisingly, DHMF yields of up to 99% could be achieved in tetrahydrofuran at a very low temperature of 40 °C over various amine−metal complexes, such as Cm*Ru(HTsDPEN), Cp*Ir(HTsDPEN), Cp*Ir(HTsDACH), and Cp*Ir(NHCPh2C6H4) (Cm* = cymene; Cp* = 1,2,3,4,5-pentamethylcyclopenta-1,3-diene), suggesting that CTH was absolutely workable for the selective hydrogenation of HMF into DHMF (Figure 5). Unsatisfactorily, homogeneous catalysts, especially Cp*Ir(HTsDACH) and Cp*Ir(NHCPh2C6H4), showed poor tolerance to formic acid; half of their initial catalytic performance was lost even in several minutes. In this situation, Pd/C was found to be capable of tolerating formic acid, and when it was used as a

Figure 6. Plausible reaction mechanism for the selective hydrogenation of HMF into DHMF over ZrO(OH)2 in the presence of ethanol. Adapted with permission from ref 203. Copyright 2016 Royal Society of Chemistry.

Enlightened by the acid−base bifunctional actions of ZrO(OH)2, Li et al. designed a series of zirconium-based organic− inorganic coordination polymers such as zirconium alkyltriphosphate nanohybrid (Zr-ATPN), zirconium benzylphosphonate nanohybrid (Zr-BPPN), and zirconium furandicarboxylate nanohybrid (Zr-FDCN) for the selective hydrogenation of HMF.204−206 In the presence of isopropanol, the DHMF yields

Figure 5. Selective hydrogenation of HMF into DHMF over Cp*Ir(HTsDPEN) in the presence of formic acid. Adapted with permission from ref 199. Copyright 2010 John Wiley and Sons. 2964

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis were as high as 98, 93, and 87% at 140, 120, and 140 °C for 2, 2, and 8 h, respectively. Gratifyingly, under similar reaction conditions, other types of catalysts, such as hydroxyapatiteencapsulated ferric oxide (Fe2O3@HAP),207 mesoporouscarbon-supported cobaltosic oxide (Co3O4@MC),208 silicon carbide-supported gold (Au/SiC),209 cobaltosic oxide-supported ruthenium (Ru/Co3 O4 ),210 and layered double hydroxide-supported ruthenium (Ru/ZnAlZr-LDH),211 also displayed superior catalytic performance, with 72, 97, 83.7, 82 and 94% yields of DHMF, respectively, clearly indicating that isopropanol possesses a broad universality toward various catalysts, so it is considered as an excellent hydrogen donor for the selective hydrogenation of HMF into DHMF, which can be mainly ascribed to its lower reduction potential.204−211 According to previous reports,212−214 the reduction potential (ΔHf°) is defined as the difference between the standard molar enthalpies of formation of an alcohol and its corresponding carbonyl compound, and it represents the complexity of hydrogen abstraction. As shown in Table 3, the reduction

Figure 7. Simplified reaction pathway for the selective hydrogenation of HMF into DHMF over MgO in the presence of methanol. Adapted with permission from ref 216. Copyright 2014 Elsevier.

formate, would further decompose into the above-mentioned gaseous coproducts, which were easily removed in the reactor depressurization process.216 In addition to the common alcohols, benzyl alcohol was also tested by Gao et al.171 in the presence of carbon-supported ruthenium and cobalt (RuCo/C), and the results revealed that it displayed a strong ability to act as a hydrogen donor for the selective hydrogenation of HMF into DHMF, giving an acceptable yield of 86.9% at 150 °C for 10 h. More recently, 1,4-butanediol, a renewable diol,217−219 was successfully employed for the selective hydrogenation of HMF. When this reaction was carried out in a continuous-flow reactor over alumina-supported copper (Cu/AlOx), a DHMF yield of 93% was achieved in only 0.6 min at 220 °C.137 Notably, one molecule of BDO can provide two molecules of H2 and can then be converted into γ-butyrolactone (GBL), a versatile intermediate for the synthesis of fine chemicals, by the oxygenfree dehydrogenation and ring-closure reaction.220−223 2.3. Electrocatalytic Hydrogenation Pathway. Electrocatalytic hydrogenation is considered as a green pathway for the selective hydrogenation of HMF because it can be performed at room temperature and atmospheric pressure in an aqueous solution, in which water and protons generally act as hydrogen donors.224−226 Therefore, the electrocatalytic hydrogenation pathway has been gradually used for the production of DHMF in recent years (Figure 8). In 2013, Kwon et al.227 first studied

Table 3. Reduction Potentials (ΔHf°) of Various Alcohols alcohol

ΔHf° (kJ/mol)

methanol ethanol propanol isopropanol 1-butanol 2-butanol

130.1a 85.4a 87.3b 70.0b 79.7b 69.3b

The numerical value of ΔH°f was calculated according to the definition of ΔHf°. bThe numerical value of ΔHf° was obtained from van der Waal et al.213 a

potentials of various alcohols decrease in the order methanol > ethanol > 1-butanol > isopropanol ≈ 2-butanol. Compared with primary alcohols, secondary alcohols, which have lower reduction potentials, are more appropriate as hydrogen donors in the process of CTH, and this may be the reason that isopropanol is widely used for the selective hydrogenation of HMF into DHMF. Because it has a similar reduction potential as isopropanol, 2-butanol should also be a superior hydrogen donor. However, to the best of our knowledge, prior to 2017 it had not been employed for the selective hydrogenation of HMF into DHMF. More recently, our group confirmed the capacity of 2-butanol. When it was used as a hydrogen donor over a magnetic zirconium hydroxide (MZH) with a moderate Zr/Fe ratio of 2, a 89.6% yield of DHMF was obtained in 5 h at 150 °C.215 Furthermore, among various alcohols, methanol has the highest reduction potential, and therefore, it does not easily act as a hydrogen donor in theory. Unexpectedly, Pasini et al.216 reported very interesting results in which methanol could be applied as an active hydrogen donor for the selective hydrogenation of HMF over the high surface area of magnesia (MgO). After 3 h at 160 °C, a DHMF yield of almost 100% was achieved, and more surprisingly, the coproducts were exclusively gaseous compounds, such as carbon monoxide (CO), carbon dioxide (CO2), and methane (CH4). From experimental results and theoretical calculations, further study of the reaction mechanism revealed that the resulting formaldehyde could react with the remaining methanol to produce the hemiacetal, which could also act as a hydrogen donor for the selective hydrogenation of HMF into DHMF (Figure 7). Meanwhile, its dehydrogenated product, methyl

Figure 8. Electrocatalytic hydrogenation of HMF into DHMF in an aqueous solution.227−229

the selective hydrogenation of HMF via an electrocatalytic pathway in 0.1 M sodium sulfate (Na2SO4). By correlation of the voltammetry with online product analysis, iron (Fe), nickel (Ni), silver (Ag), zinc (Zn), cadmium (Cd), and indium (In) were verified to be effective for the formation of DHMF. Among these electrodes, Ag showed the best catalytic 2965

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis performance, leading to 85% selectivity at −0.81 V. However, when the selective hydrogenation of HMF was conducted in 0.5 M sulfuric acid (H2SO4),228 the most effective electrode was Ni, indicating that the electrocatalytic hydrogenation of HMF into DHMF was strongly influenced by the types of electrodes and electrolytes, as was further confirmed by the results of Nilges and Schröder.225 In addition, the onset potentials of the target product formation and hydrogen evolution reaction (HER) in the electrocatalytic hydrogenation of HMF into DHMF are closely correlated to the pH of the electrolyte,227−229 as shown in Figure 9. In a neutral solution (pH 7),227 the onset potentials of

onset potentials of the HER. In both neutral and acidic solutions, the HER could occur via the generation of Hads, regardless of whether its source was H2O (H2O + e− ⇄ Hads + OH−) or H+ (H+ + e− ⇄ Hads). Interestingly, the HER was much more effective in the acidic solution than in the neutral solution, suggesting that Hads on the metal surface was involved in the formation of the target product in the acidic solution (HMF + 2Hads ⇄ DHMF).228 According to the previous studies, a plausible mechanism for the electrocatalytic hydrogenation of HMF into DHMF on the metal electrodes was proposed by Choi and co-workers.230 As shown in Figure 10, the formation of DHMF could proceed through transfer of e− to HMF to produce an anionic intermediate that further reacted with H+ or through the simultaneous transfer of e− and H+ to HMF (blue arrows in Figure 10). On the other hand, the HER could result in the formation of Hads on the metal electrode surface, which would further drive the selective hydrogenation of HMF into DHMF (red arrows in Figure 10). Furthermore, Choi and co-workers also constructed an ideal photoelectrocatalytic pathway simply by replacing a metal anode (e.g., Pt) in the electrocatalytic pathway with a photoanode (e.g., BiVO4).230 Under illumination, BiVO4 generated electron−hole pairs. The holes were used for water splitting, while the photoexcited electrons were transferred to the Ag cathode for the selective hydrogenation of HMF (Figure 11), giving satisfactory results with 95% DHMF

Figure 9. Comparison of the onset potentials of the product formation and the HER on (a) transition metals and (b) post-transition metals in neutral and acidic solutions. The onset potentials for the HER were obtained at a current density of −0.5 mA·cm−2. Adapted with permission from ref 228. Copyright 2015 John Wiley and Sons.

the target product formation on all of the metal electrodes were −0.5 ± 0.2 V, which are less negative than those of the HER, which ranged from −1 to −0.4 V on transition metals and −1.5 to −1 V on post-transition metals. These similar onset potentials implied a weak catalytic effect on the electron transfer reaction, which is the first step in the HER. Namely, the adsorbed hydrogen (Hads) on the metal surface was unlikely to be involved in the formation of the target product, and therefore, this process should occur directly through water molecules from the neutral solution: HMF + 2H2O + 2e− ⇄ DHMF + 2OH−.227 However, in an acidic solution (pH < 7),228 the onset potentials of the target product formation shifted to more positive values, ranging from −0.4 to 0 V on transition metals and −0.7 to −0.25 V on post-transition metals. Moreover, the onset potentials of the target product formation were also found to be greatly associated with the

Figure 11. Comparison beween (a) the photoelectrocatalytic pathway and (b) the electrocatalytic pathway for the selective hydrogenation of HMF into DHMF. Adapted from ref 230. Copyright 2016 American Chemical Society.

selectivity and 94% Faradaic efficiency (FE).230 This photoelectrocatalytic pathway provides a practical, inexpensive, and

Figure 10. Plausible reaction mechanism for the selective hydrogenation of HMF into DHMF over a Ag electrode via an electrocatalytic pathway. Adapted from ref 230. Copyright 2016 American Chemical Society. 2966

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

this is a sustainable pathway for the conversion of biomass using solar radiation as a driving force. However, it should be emphatically noted that the catalytic performance of this pathway is not ideal, and therefore, it is far from large-scale applications. On this occasion, how to design and prepare a more effective photocatalyst and balance the roles of the photocatalyst in the production and subsequent activation of hydrogen should be the research priority for the photocatalytic hydrogenation of HMF into DHMF. 2.5. Disproportionation Reaction Pathway. Except for the above-mentioned pathways, the selective hydrogenation of HMF into DHMF can also be accomplished by the Cannizzaro reaction,232−234 which is a base-induced disproportionation reaction between the same aldehyde compound lacking a hydrogen atom at the α-position,235−237 where the aldehyde compound serves simultaneously as a hydrogen donor and a hydrogen acceptor. However, for many decades, only two reports on the Cannizzaro reaction of HMF were published, by Blanksma in 1910238 and by Middendorp in 1919.239 Excitingly, Kang et al.240 in 2012 and Subbiah et al.241 in 2013 restudied the Cannizzaro reaction of HMF. More importantly, outstanding DHMF yields of 100 and 90% could be achieved at 0 °C for 1 and 4 h over sodium hydroxide (NaOH) and sodium hydride (NaH) in 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([EMIM]TFSI) and tetrahydrofuran, respectively. At the same time, 5-hydroxymethylfuranoic acid (HMFA), a corresponding oxidative product, was also obtained in high yields of 84 and 89%. In theory, the Cannizzaro reaction can form an equimolar mixture of the oxidized and reduced products. If the target product is one of the products, the application of the Cannizzaro reaction will be heavily restrained. However, if the Cannizzaro reaction is applied to HMF (Figure 13), it should be an atom-economic pathway because DHMF and HMFA, the equally important products of HMF, can be simultaneously synthesized.

effective approach for decreasing the external voltage necessary for the selective hydrogenation of HMF into DHMF, which is an excellent example of coupling solar energy conversion and biomass conversion. It should be noted that although the electrocatalytic pathway has been deeply studied, its large-scale application in the selective hydrogenation of HMF into DHMF is currently limited. One of the reasons for this is that the electrocatalytic pathway is a multidisciplinary approach, and for its successful implementation, wonderful cooperation among electrochemistry, electrochemical engineering, organic chemistry, catalysis, and process design is indispensable. Unfortunately, this cooperation has not been successfully implemented until now. Furthermore, the electrocatalytic pathway requires a high input of electricity, and considering current electricity prices, this pathway still lacks economic superiority. In this aspect, the photoelectrocatalytic pathway should be more competitive because it can largely reduce the need for external electricity (Figure 11). If a more effective and inexpensive photoanode with a smaller band gap is developed and a more appropriate cell with a minimum internal resistance drop is designed, the photoelectrocatalytic hydrogenation of HMF into DHMF will probably be achieved with no external electricity input, in which the total energy efficiency should also be considered and compared between the photoelectrocatalytic pathway and the electrocatalytic pathway plus a single solar cell. 2.4. Photocatalytic Hydrogenation Pathway. Inspired by the photoelectrocatalytic pathway, a pure photocatalytic pathway was developed by Guo and Chen231 in 2016 for the selective hydrogenation of HMF into DHMF, in which graphitic carbon nitride-supported platinum (Pt/g-C3N4) was selected as a multifunctional photocatalyst and triethylamine (TEA) was employed as a sacrificial electron donor in aqueous solution (Figure 12). The experimental results indicated that

Figure 13. Selective hydrogenation of HMF into DHMF over NaOH or NaH via the Cannizzaro reaction. Adapted with permission from ref 240. Copyright 2012 Elsevier.

More recently, inspired by the Cannizzaro reaction of HMF, a novel acid-catalyzed disproportionation reaction over trimethylaluminum (AlMe3) in acetonitrile was developed for the selective hydrogenation of HMF into DHMF by Li et al.242 In this reaction, HMF itself acted as both a hydrogen donor and a hydrogen acceptor. Differently, the corresponding oxidative product was 2,5-diformylfuran (DFF) rather than HMFA because the nature of the acid-catalyzed disproportionation reaction is the intermolecular MPV reaction. In addition, although DHMF and DFF were successfully formed in the presence of AlMe3, the HMF conversion was only 26.7% at 80 °C for 1 h. By a combination of the theoretical and experimental results, the lower conversion of HMF was proved to be due to the redox equilibrium in HMF and DHMF/DFF (Figure 14). Fortunately, DHMF shows poor solubility in acetonitrile, and therefore, its precipitation from the reaction mixture should shift the redox equilibrium toward the desired products. On the basis of this speculation, a cooling treatment

Figure 12. Selective hydrogenation of HMF into DHMF via a photocatalytic pathway in water over Pt/g-C3N4 with TEA as a sacrificial electron donor. Adapted with permission from ref 231. Copyright 2016 Royal Society of Chemistry.

Pt/g-C3N4 could not only facilitate photoinduced water splitting to produce hydrogen but also promote the subsequent activation of the produced hydrogen for the selective hydrogenation of HMF as a result of the synergistic action of Pt and g-C3N4.231 Under visible-light irradiation using a 210 W xenon lamp equipped with a 420 nm cutoff filter, a 6.5% yield of DHMF was observed at 80 °C for 4 h.231 To the best of our knowledge, this is the first report of the direct photocatalytic hydrogenation of HMF to give DHMF, and most importantly, 2967

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

Figure 15. Selective hydrogenation of HMF into DHMF via a biocatalytic pathway in the presence of Escherichia coli CCZU-K14.254

in an ice−water bath at an interval of 10 min was applied in the acid-catalyzed disproportionation reaction of HMF. Surprisingly, with this strategy, the HMF conversion obviously increased by 71.3% under the same reaction conditions, and the relative content of DFF was much higher than that of DHMF in the supernatant liquid. Meanwhile, the relative content of DHMF was much higher than that of DFF in the precipitate,242 indicating that the difference in solubility of DHMF and DFF in acetonitrile not only permitted the redox equilibrium to shift but also provided a promising protocol for the crude separation of DHMF from DFF, which was beneficial for the production and purification of DHMF.

as 100 and 90.2%, respectively, at the same reaction temperature. Obviously, the biocatalytic hydrogenation of HMF into DHMF was successfully realized in the presence of E. coli CCZU-K14. Moreover, the use of reduced nicotinamide adenine dinucleotide (NADH) or reduced nicotinamide adenine dinucleotide phosphate (NADPH) as a cofactor was crucial to the biocatalytic hydrogenation reaction, and its effective regeneration was closely related to the cosubstrate glucose and its concentration. Thus, the appropriate concentration of glucose was generally added in the selective hydrogenation of HMF into DHMF via the biocatalytic pathway.252−254 Nevertheless, the reported biocatalytic pathways, although promising, must still be validated under industrial conditions. For industrial production, immobilized whole cells with excellent reusability and high productivity are highly recommended.

3. BIOCATALYTIC HYDROGENATION OF HMF INTO DHMF As can be seen from the above descriptions, the chemocatalytic pathway is the main method for the selective hydrogenation of HMF into DHMF, and it has therefore been widely studied in recent years. In contrast to the chemocatalytic pathway, the biocatalytic pathway, which has a number of advantages such as excellent selectivity, high efficiency, environmental friendliness, and mild reaction conditions, may be a more attractive option.243−245 Unfortunately, it receives little attention, which may be ascribed to the high cost and vulnerability of enzymes. Compared with isolated enzymes, whole cells are preferable for the selective hydrogenation of HMF into DHMF because they are not only inexpensive and stable but also do not require complex systems for the separation and purification of enzymes.246−248 However, the production of DHMF from HMF using the whole cells is extremely challenging because HMF is a well-known inhibitor and toxic compound to microorganisms.249−251 Hence, searching for a new microbial strain with high tolerance to HMF is essential for the biocatalytic hydrogenation of HMF into DHMF. In 2004, Saccharomyces cerevisiae ATCC 211239 and NRRL Y-12632 and Pichia stipitis NRRL Y-7124 were evaluated by Liu et al.252 to examine their responses to HMF. Surprisingly, among the three strains, S. cerevisiae NRRL Y-12632 not only could tolerate 30 mM HMF but also could convert it into DHMF. As far as we know, this is the first time that DHMF was identified as the final hydrogenation product of HMF by yeast. Subsequently, Meyerozyma guilliermondi SC1103 was isolated by Li et al.253 in 2017, and it was proved to be more effective for the biocatalytic hydrogenation of HMF into DHMF. When the concentration of HMF was 100 mM, a DHMF yield of 86% could be achieved at 35 °C for 12 h. Following this exciting result, He et al.254 found that recombinational Escherichia coli CCZU-K14 also displayed an amazing tolerance and catalytic ability toward HMF (Figure 15). When the concentration of HMF was increased to 400 mM, a 70.2% yield of DHMF was still obtained at 30 °C for 72 h. When 100 and 200 mM HMF were used as substrates, the DHMF yield could even be as high

4. FURTHER TRANSFORMATION OF DHMF INTO VALUABLE DERIVATIVES DHMF has a peculiar symmetrical structure containing a furan ring and two hydroxyl groups, determining its wide potential applications. It is well-known that DHMF is an important intermediate in the total hydrogenation and hydrogenolysis of HMF, and therefore, by adoption of appropriate reaction conditions, it can be further transformed into DMF, DHMTHF, DMTHF, HTO, HDO, HHD, or HMCPN, which have been involved in previous reviews.5,214,255,256 Moreover, the earlier studies showed that DHMF can also be used for the synthesis of DFF,257 2,5-furandicarboxylic acid (FDCA),258−261 and various crown ethers.111 However, these applications have been rarely reported in the present studies. Thus, in the following section, only a few typical applications of DHMF will be introduced in detail. 4.1. Through the Etherification Reaction. Because of its higher energy density, higher cetane number, higher miscibility, and stronger stability, 2,5-bis(alkoxymethyl)furan (BAMF) has been assessed as a more attractive diesel additive than 5alkoxymethylfurfural (AMF).262−267 Similar to the preparation of AMF from HMF,268−278 BAMF can be synthesized by the etherification of DHMF with the corresponding alcohols in the presence of acid catalysts (Table 4). For instance, Balakrishnan et al.112 first reported the etherification of DHMF with ethanol in 2012. The results demonstrated that among various acid catalysts, such as H2SO4, Amberlyst-15, Amberlite IR120, Dowex 50WX8, and Dowex DR2030, Amberlyst-15 showed an outstanding catalytic performance with an 80% yield of 2,5bis(ethoxymethyl)furan (BEMF) at a low reaction temperature of 40 °C for 16 h, which was in accordance with the results of Sacia et al.262 More importantly, Amberlyst-15 could also be employed for the efficient preparation of 2,5-bis(propoxymethyl)furan (BPMF)263 and 2,5-bis(butoxymethyl)furan (BBMF)263 via the etherification of DHMF with propanol and 1-butanol, respectively. However, if methanol was applied for the etherification of DHMF, the catalytic performance of Amberlyst-15 was not satisfactory, and the yield of 2,5-bis(methoxymethyl)furan (BMMF) was only 57% at 60

Figure 14. Selective hydrogenation of HMF into DHMF over AlMe3 via the MPV reaction. Adapted with permission from ref 242. Copyright 2017 John Wiley and Sons.

2968

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis Table 4. Etherification of DHMF into BAMF in Various Alcohols alcohol methanol methanol methanol methanol methanol methanol methanol ethanol ethanol ethanol ethanol ethanol ethanol ethanol propanol propanol propanol isopropanol isopropanol isopropanol isopropanol 1-butanol 1-butanol 1-butanol 1-butanol 2-butanol 2-butanol 2-butanol

catalyst HZSM-5 (Si/Al HZSM-5 (Si/Al HZSM-5 (Si/Al HZSM-5 (Si/Al HZSM-5 (Si/Al Amberlyst-15 Amberlyst-15 Amberlyst-15 Amberlyst-15 Amberlyst-15 Amberlyst-15 Amberlyst-15 Sn-Beta Hf-Beta Amberlyst-15 Amberlyst-15 Sn-Beta Sn-Beta Zr-Mont Sn-Beta Sn-Beta Amberlyst-15 Amberlyst-15 Zr-Mont Sn-Beta Zr-Mont Sn-Beta Hf-Beta

= = = = =

25) 38) 300) 25) 25)

temperature (°C)

time (h)

DHMF conversion (%)

BAMF yield (%)

ref

100 100 100 120 140 60 60 40 60 60 60 60 180 120 60 60 180 180 150 180 180 60 60 150 180 150 180 120

3 3 3 12 8 10 10 16 5 18 10 10 6 24 10 10 6 3 1 3 6 10 10 3 6 1 6 1

100 100 100 − − 99 − − − − 99 − − − 99 − − − − 100 − 99 − − − − − −

70 68 69 68a 59a 57 50a 80 74 64a 70 70a 68.1a 67a 74 72a 61.3a 82.5a 95a 85.6 79.6a 74 71a 49a 60.5a 96a 73.1a 81a

113 113 113 113 113 263 263 112 112 112 263 263 266 267 263 263 266 264 265 266 266 263 263 265 266 265 266 267

HMF was used as a starting substrate, and it was first converted into DHMF over the appropriate hydrogenation catalyst in the preparation process of BAMF. a

Figure 16. Direct preparation of BAMF from HMF in various alcohols.262−267

°C for 10 h, which might be attributed to the high polarity of methanol, leading to the formation of many byproducts.263 Given this phenomenon, Cao et al.113 selected acidic zeolites to obtain a satisfactory yield of BMMF. Among various acidic zeolites, HZSM-5 with a Si/Al ratio of 25 was the most effective

for the etherification of DHMF with methanol. After 3 h at 100 °C, a BMMF yield of 70% could be achieved, which was higher than that over HZSM-5 with a Si/Al ratio of 38 or 300. It should be noted that a smaller Si/Al ratio of the zeolite means more acidic sites and stronger hydrophilicities, which facilitated 2969

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

controlling the number and concentration of M2. With this strategy, PFS/M2 displayed excellent multishape memory and recovery behavior (Figure 18). Subsequently, an ultraviolet-

the etherification reaction between DHMF and methanol on the surface of the catalyst.113 On the basis of these results, it can be seen that the properties of the catalyst and the type of alcohol have a great effect on the etherification of DHMF into BAMF. In addition, it should be especially noted that various BAMFs could also be directly produced from HMF in the corresponding alcohols (Table 4),262−267 in which two successive steps were involved in the same reactor: the selective hydrogenation of HMF into DHMF over a hydrogenation catalyst and the subsequent etherification of DHMF into BAMF over an acid catalyst (Figure 16). More gratifyingly, an obvious loss in the yield of BAMF was not observed. Therefore, this one-pot, two-step process was more desirable for the preparation of BAMF from the viewpoint of practical production. 4.2. Through the Polymerization Reaction. FDCA has been widely considered as a suitable replacement for terephthalic acid in the synthesis of various polymers.279−283 Correspondingly, it is reasonable to assume that DHMF is also a suitable replacement for terephthalyl alcohol.120 In the early application of DHMF, it was commonly used as a comonomer for the preparation of furan-based polyisocyanurates, polyurethanes, polyacrylonitriles, and polyethers, which were the corresponding components of foams and fibers.284−291 Compared with the benzene ring, the furan ring in DHMF is more reactive and less stable,120 and therefore, melt polymerization was rarely employed for the polymerization of DHMF with other substrates because it required a very high melt temperature, which was not favorable for the reduced furan nuclei and led to the occurrence of undesired reactions.284 Therefore, solution polymerization should be more feasible. Recently, Yoshie and co-workers reported a series of studies292−295 in which poly(2,5-furandimethylene succinate) (PFS) was prepared by the polymerization of DHMF with succinic acid (SA) under basic conditions. Interestingly, this copolymer could be further cross-linked with bismaleimide (M2) by the Diels−Alder reaction to form a network polymer (PFS/M2) (Figure 17), and meanwhile, its mechanical and healing properties could also be correspondingly adjusted by

Figure 18. (a) Preparation process, (b) shape memory, and (c) shape recovery of PFS/M2. Adapted from ref 295. Copyright 2014 American Chemical Society.

induced photopolymerization method was developed by Jang et al.118 for the preparation of a sequence of novel network polymers from vegetable oils, such as acrylated epoxidized soybean oil (AESO), acrylated castor oil (ACO), and acrylated 7,10-dihydroxy-(8E)-octadecenoic acid (ADOD), in which 2,5furan diacrylate, which was synthesized by the acylation of DHMF with acryloyl chloride, was used as a difunctional stiffener, and its addition could increase the tensile strength of the network polymers by up to 1.4−4.2 times relative to those obtained without the addition. It is worth pointing out that photopolymerization was very effective, and after only 15 min, the gel content in all of the network polymers was higher than 97%.118 More recently, Jiang et al.296 explored an enzymatic method for the polymerization of DHMF with various diacid ethyl esters such as diethyl succinate, diethyl glutarate, diethyl adipate, diethyl suberate, diethyl sebacate, and diethyl dodecanedioate in the presence of Candida antarctica lipase B (CALB). Through a three-stage process, six polyesters were prepared (Figure 19), and more importantly, they had diverse functions as self-healing materials, biomedical materials, and degradable materials. Although enzymatic polymerization has been intensively used for the synthesis of various bio-based polyesters from renewable resources, such as succinic acid, fatty acid, glycerol, and 1,4-butanediol,297−303 this method has not

Figure 17. Synthesis of PFS/M2 via the polymerization of DHMF and SA with M2. Adapted with permission from ref 293. Copyright 2013 Elsevier. 2970

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

external molecular H2 should be the most likely for the industrial production of DHMF. Unfortunately, this process has not yet been realized until now despite the achievement of a series of breakthroughs. Among all of the possible bottlenecks, the main one is the high cost of HMF. If HMF is prepared in an inexpensive way, greater progress will be readily achieved in this research area; hence, considerable attention should be persistently focused on the economical synthesis of HMF from carbohydrates or biomass. Moreover, to promote the industrial production and further application of DHMF, the following potential issues should also be further reinforced in future studies: (I) Systematic construction of exclusive catalysts. Because of the special structure and reactivity of DHMF, catalysts must be extremely selective in their production and application. Considering sustainability, recyclability, and environmental friendliness, heterogeneous catalysts with nonprecious metals, specific sizes, high dispersity, and excellent stability are particularly desired. In addition, if DHMF is directly produced and applied via a one-pot reaction from carbohydrates or biomass, the tolerance of the catalysts toward the complex systems should also be considered. (II) Intensive exploration of appropriate reaction media. During the production and application of DHMF, the reaction medium can act not only as the solvent but also as a hydrogen donor and reactant. From the viewpoint of green chemistry, the development of inexpensive and renewable reaction media with strong solvencies and particular functionalities is imminently necessary. (III) Comprehensive optimization of reaction parameters. Once the exclusive catalysts and reaction media are chosen in the production and application of DHMF, various reaction parameters such as reaction temperature, reaction time, catalyst amount, and substrate concentration will have a more direct influence on the formation of the target products. Thus, to get higher yields, they should be thoroughly optimized. (IV) Strategic establishment of separation methods. It is generally known that DHMF is not very stable, especially in high-temperature circumstances, and the reaction mixture is quite complicated in its production and application. In this situation, the effective separation of the target products from the complicated reaction mixture is the key for their practical uses. Therefore, in addition to getting higher yields of the target products, establishing high-efficiency, energy-efficient separation methods based on the respective physicochemical properties of various target products is urgently needed. (V) Innovative design of continuous reactors. Whether DHMF is produced or applied, multifunctional continuous reactors including catalytic equipment for the production and application of DHMF, recovery equipment for recycling of

Figure 19. Enzymatic polymerization of DHMF with various diacid ethyl esters over CALB. Adapted from ref 296. Copyright 2014 American Chemical Society.

received enough attention in the polymerization of DHMF; hence, its application should be reinforced in the future. 4.3. Through the Rearrangement Reaction. In 2016, α6-hydroxy-6-methyl-4-enyl-2H-pyran-3-one (HMEPO), which is a novel and attractive pyranone for the synthesis of sugar analogues and compounds with excellent biological activities,304−307 was creatively prepared by Gelmini et al.114 via ring rearrangement of DHMF in the presence of water over a heterogeneous acid catalyst of Amberlyst-15. Notably, the same treatment in alcohol led to etherification of the hydroxy groups, and no rearrangement reaction was observed. Under nitrogen (N2) at atmospheric pressure, the 50% yield and 70% selectivity of HMEPO could be achieved in 30 min at a moderate reaction temperature of 70 °C. Furthermore, this ring rearrangement reaction does not require the addition of oxidants, in contrast to the Achmatowicz rearrangement reaction, which is a typical oxidative process for the conversion of 2-(α-hydroxyalkyl)furans into 6-hydroxy-3(2H)-pyranones. On the basis of the previous reports308−312 and the results of gas chromatography− mass spectrometry and 1H NMR analyses, a plausible reaction mechanism for the ring rearrangement of DHMF into HMEPO involving hydration, dehydration, ring opening, keto−enol tautomerism, and ring closure is proposed in Figure 20. In addition, if the proper reaction conditions were adopted, for example, a prolonged reaction time of 5 h or the use of Amberlyst-70 as a catalyst, other interesting products, such as 1,3-dihydroxyhexan-2,5-dione and 1,4-dihydroxyhexan-2,5dione, could also be preferentially formed from the parallel hydration of the ketonic intermediate or from the consecutive ring opening and hydration of HMEPO. Conclusively, this mechanistic investigation opens an avenue for the development of a general method for the transformation of various dihydroxyfurans.

5. CONCLUSIONS AND PERSPECTIVES As mentioned above, DHMF is a very promising versatile diol, and it can be produced from the selective hydrogenation of biomass-derived HMF via various chemocatalytic and biocatalytic pathways. From the current point of view, among these pathways the conventional hydrogenation pathway with

Figure 20. Plausible reaction mechanism for the ring rearrangement of DHMF into HMEPO over Amberlyst-15 in the presence of water. Adapted with permission from ref 114. Copyright 2016 Elsevier. 2971

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

(12) Farran, A.; Cai, C.; Sandoval, M.; Xu, Y. M.; Liu, J.; Hernaiz, M. J.; Linhardt, R. J. Green solvents in carbohydrate Chemistry: From raw materials to fine chemicals. Chem. Rev. 2015, 115, 6811−6853. (13) Yan, K.; Jarvis, C.; Gu, J.; Yan, Y. Production and catalytic transformation of levulinic acid: A platform for speciality chemicals and fuels. Renewable Sustainable Energy Rev. 2015, 51, 986−997. (14) Hu, L.; Lin, L.; Wu, Z.; Zhou, S. Y.; Liu, S. J. Chemocatalytic hydrolysis of cellulose into glucose over solid acid catalysts. Appl. Catal., B 2015, 174−175, 225−243. (15) Li, X. D.; Jia, P.; Wang, T. F. Furfural: A promising platform compound for sustainable production of C4 and C5 Chemicals. ACS Catal. 2016, 6, 7621−7640. (16) Kobayashi, H.; Kaiki, H.; Shrotri, A.; Techikawara, K.; Fukuoka, A. Hydrolysis of woody biomass by a biomass-derived reusable heterogeneous catalyst. Chem. Sci. 2016, 7, 692−696. (17) Pileidis, F. D.; Titirici, M. M. Levulinic acid biorefineries: New challenges for efficient utilization of biomass. ChemSusChem 2016, 9, 562−582. (18) Esposito, D.; Antonietti, M. Redefining biorefinery: The search for unconventional building blocks for materials. Chem. Soc. Rev. 2015, 44, 5821−5835. (19) Liu, D. J.; Chen, E. Y. X. Organocatalysis in biorefining for biomass conversion and upgrading. Green Chem. 2014, 16, 964−981. (20) Karinen, R.; Vilonen, K.; Niemela, M. Biorefining: Heterogeneously catalyzed reactions of carbohydrates for the production of furfural and hydroxymethylfurfural. ChemSusChem 2011, 4, 1002− 1016. (21) Román-Leshkov, Y.; Chheda, J. N.; Dumesic, J. A. Phase modifiers promote efficient production of hydroxymethylfurfural from fructose. Science 2006, 312, 1933−1937. (22) Qi, X. H.; Watanabe, M.; Aida, T. M.; Smith, R. L., Jr. Efficient process for conversion of fructose to 5-hydroxymethylfurfural with ionic liquids. Green Chem. 2009, 11, 1327−1331. (23) Chen, J. Z.; Li, K. G.; Chen, L. M.; Liu, R. L.; Huang, X.; Ye, D. Q. Conversion of fructose into 5-hydroxymethylfurfural catalyzed by recyclable sulfonic acid-functionalized metal-organic frameworks. Green Chem. 2014, 16, 2490−2499. (24) Jadhav, A. H.; Chinnappan, A.; Patil, R. H.; Kostjuk, S. V.; Kim, H. Green chemical conversion of fructose into 5-hydroxymethylfurfural (HMF) using unsymmetrical dicationic ionic liquids under mild reaction condition. Chem. Eng. J. 2014, 243, 92−98. (25) Wang, H. L.; Kong, Q. Q.; Wang, Y. X.; Deng, T. S.; Chen, C. M.; Hou, X. L.; Zhu, Y. L. Graphene oxide catalyzed dehydration of fructose into 5-hydroxymethylfurfural with isopropanol as cosolvent. ChemCatChem 2014, 6, 728−732. (26) Hu, Z. G.; Peng, Y. W.; Gao, Y. J.; Qian, Y. H.; Ying, S. M.; Yuan, D. Q.; Horike, S.; Ogiwara, N.; Babarao, R.; Wang, Y. X.; Yan, N.; Zhao, D. Direct synthesis of hierarchically porous metal-organic frameworks with high stability and strong Brønsted acidity: The decisive role of hafnium in efficient and selective fructose dehydration. Chem. Mater. 2016, 28, 2659−2667. (27) Zhao, H. B.; Holladay, J. E.; Brown, H.; Zhang, Z. C. Metal chlorides in ionic liquid solvents convert sugars to 5-hydroxymethylfurfural. Science 2007, 316, 1597−1600. (28) Yong, G.; Zhang, Y.; Ying, J. Y. Efficient catalytic system for the selective production of 5-hydroxymethylfurfural from glucose and fructose. Angew. Chem., Int. Ed. 2008, 47, 9345−9348. (29) Hu, S. Q.; Zhang, Z. F.; Song, J. L.; Zhou, Y. X.; Han, B. X. Efficient conversion of glucose into 5-hydroxymethylfurfural catalyzed by a common Lewis acid SnCl4 in an ionic liquid. Green Chem. 2009, 11, 1746−1749. (30) 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 Bronsted acid catalysts in glucose and fructose conversion to 5-(hydroxymethyl)furfural and levulinic acid in aqueous media. J. Am. Chem. Soc. 2013, 135, 3997− 4006. (31) Jiménez-Morales, I.; Teckchandani-Ortiz, A.; SantamaríaGonzález, J.; Maireles-Torres, P.; Jiménez-López, A. Selective

catalysts and reaction media, and separation equipment for isolation of the target products should be creatively designed according to the corresponding catalytic systems. Apart from these scientific issues, in-depth techno-economic analyses of various catalytic pathways should also be intensively performed to appraise their availability for the industrial production and application of DHMF. All in all, we must continue our efforts in the development of appropriate approaches to reduce the synthesis costs of HMF and increase the application economy of DHMF. Despite facing numerous difficulties and challenges, we still believe that the future prospects are exceptionally bright.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Telephone/Fax: +86-051783526983. ORCID

Lei Hu: 0000-0002-0547-6265 Xing Tang: 0000-0003-3428-776X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Natural Science Foundation of China (21506071), the Natural Science Foundation of Jiangsu Province (BK20150413), and the Natural Science Foundation of the Higher Education Institutions of Jiangsu Province (16KJA530001).



REFERENCES

(1) Binder, J. B.; Raines, R. T. Simple chemical transformation of lignocellulosic biomass into furans for fuels and chemicals. J. Am. Chem. Soc. 2009, 131, 1979−1985. (2) Zhou, C. H.; Xia, X.; Lin, C. X.; Tong, D. S.; Beltramini, J. Catalytic conversion of lignocellulosic biomass to fine chemicals and fuels. Chem. Soc. Rev. 2011, 40, 5588−5617. (3) Chatterjee, C.; Pong, F.; Sen, A. Chemical conversion pathways for carbohydrates. Green Chem. 2015, 17, 40−71. (4) Zhang, Z. R.; Song, J. L.; Han, B. X. Catalytic transformation of lignocellulose into chemicals and fuel products in ionic liquids. Chem. Rev. 2017, 117, 6834−6880. (5) Hu, L.; Lin, L.; Wu, Z.; Zhou, S. Y.; Liu, S. J. Recent advances in catalytic transformation of biomass-derived 5-hydroxymethylfurfural into the innovative fuels and chemicals. Renewable Sustainable Energy Rev. 2017, 74, 230−257. (6) Hu, L.; Zhao, G.; Hao, W. W.; Tang, X.; Sun, Y.; Lin, L.; Liu, S. J. Catalytic conversion of biomass-derived carbohydrates into fuels and chemicals via furanic aldehydes. RSC Adv. 2012, 2, 11184−11206. (7) Xu, C. P.; Arancon, R. A. D.; Labidi, J.; Luque, R. Lignin depolymerisation strategies: Towards valuable chemicals and fuels. Chem. Soc. Rev. 2014, 43, 7485−7500. (8) Wang, J. J.; Xi, J. X.; Wang, Y. Q. Recent advances in the catalytic production of glucose from lignocellulosic biomass. Green Chem. 2015, 17, 737−751. (9) Zhang, Z. H.; Deng, K. J. Recent advances in the catalytic synthesis of 2,5-furandicarboxylic acid and its derivatives. ACS Catal. 2015, 5, 6529−6544. (10) Liu, B.; Zhang, Z. H. Catalytic conversion of biomass into chemicals and fuels over magnetic catalysts. ACS Catal. 2016, 6, 326− 338. (11) Li, C. Z.; Zhao, X. C.; Wang, A. Q.; Huber, G. W.; Zhang, T. Catalytic transformation of lignin for the production of chemicals and fuels. Chem. Rev. 2015, 115, 11559−11624. 2972

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

aluminium chloride catalyst in water. Green Chem. 2011, 13, 2859− 2868. (50) Nikolla, E.; Román-Leshkov, Y.; Moliner, M.; Davis, M. E. Onepot” synthesis of 5-(hydroxymethyl)furfural from carbohydrates using tin-beta zeolite. ACS Catal. 2011, 1, 408−410. (51) Yang, Y.; Xiang, X.; Tong, D. M.; Hu, C. W.; Abu-Omar, M. M. One-pot synthesis of 5-hydroxymethylfurfural directly from starch over SO42−/ZrO2-Al2O3 solid catalyst. Bioresour. Technol. 2012, 116, 302− 306. (52) Yepez, A.; Garcia, A.; Climent, M. S.; Romero, A. A.; Luque, R. Catalytic conversion of starch into valuable furan derivatives using supported metal nanoparticles on mesoporous aluminosilicate materials. Catal. Sci. Technol. 2014, 4, 428−434. (53) Roy Goswami, S.; Dumont, M. J.; Raghavan, V. Microwave assisted synthesis of 5-hydroxymethylfurfural from starch in AlCl3· 6H2O/DMSO/[BMIM]Cl system. Ind. Eng. Chem. Res. 2016, 55, 4473−4481. (54) Yu, I. K. M.; Tsang, D. C. W.; Yip, A. C. K.; Chen, S. S.; Wang, L.; Ok, Y. S.; Poon, C. S. Catalytic valorization of starch-rich food waste into hydroxymethylfurfural (HMF): Controlling relative kinetics for high productivity. Bioresour. Technol. 2017, 237, 222−230. (55) Dutta, S.; De, S.; Alam, M. I.; Abu-Omar, M. M.; Saha, B. Direct conversion of cellulose and lignocellulosic biomass into chemicals and biofuel with metal chloride catalysts. J. Catal. 2012, 288, 8−15. (56) Liu, B.; Zhang, Z. H.; Zhao, Z. B. Microwave-assisted catalytic conversion of cellulose into 5-hydroxymethylfurfural in ionic liquids. Chem. Eng. J. 2013, 215−216, 517−521. (57) Xiao, S. H.; Liu, B.; Wang, Y. M.; Fang, Z. F.; Zhang, Z. H. Efficient conversion of cellulose into biofuel precursor 5-hydroxymethylfurfural in dimethyl sulfoxide-ionic liquid mixtures. Bioresour. Technol. 2014, 151, 361−366. (58) Zhang, X. Y.; Zhang, D.; Sun, Z.; Xue, L. F.; Wang, X. H.; Jiang, Z. J. Highly efficient preparation of HMF from cellulose using temperature-responsive heteropolyacid catalysts in cascade reaction. Appl. Catal., B 2016, 196, 50−56. (59) Yuan, B.; Guan, J.; Peng, J.; Zhu, G. Z.; Jiang, J. H. Green hydrolysis of corncob cellulose into 5-hydroxymethylfurfural using hydrophobic imidazole ionic liquids with a recyclable, magnetic metalloporphyrin catalyst. Chem. Eng. J. 2017, 330, 109−119. (60) Nguyen, T. D.; Nguyen, H. D.; Nguyen, P. T.; Nguyen, H. D. Magnetic poly(vinylsulfonic-co-divinylbenzene) catalysts for direct conversion of cellulose into 5-hydroxymethylfurfural using ionic liquids. Mater. Trans. 2015, 56, 1434−1440. (61) Zhou, P.; Zhang, Z. H. One-pot catalytic conversion of carbohydrates into furfural and 5-hydroxymethylfurfural. Catal. Sci. Technol. 2016, 6, 3694−3712. (62) Rout, P. K.; Nannaware, A. D.; Prakash, O.; Kalra, A.; Rajasekharan, R. Synthesis of hydroxymethylfurfural from cellulose using green processes: A promising biochemical and biofuel feedstock. Chem. Eng. Sci. 2016, 142, 318−346. (63) van Putten, R. J.; van der Waal, J. C.; de Jong, E.; Rasrendra, C. B.; Heeres, H. J.; de Vries, J. G. Hydroxymethylfurfural, a versatile platform chemical made from renewable resources. Chem. Rev. 2013, 113, 1499−1597. (64) Zakrzewska, M. E.; Bogel-Łukasik, E.; Bogel-Łukasik, R. Ionic liquid-mediated formation of 5-hydroxymethylfurfurals: A promising biomass-derived building block. Chem. Rev. 2011, 111, 397−417. (65) Teong, S. P.; Yi, G. S.; Zhang, Y. G. Hydroxymethylfurfural production from bioresources: Past, present and future. Green Chem. 2014, 16, 2015−2026. (66) Román-Leshkov, Y.; Barrett, C. J.; Liu, Z. Y.; Dumesic, J. A. Production of dimethylfuran for liquid fuels from biomass-derived carbohydrates. Nature 2007, 447, 982−985. (67) Verevkin, S. P.; Emel’yanenko, V. N.; Stepurko, E. N.; Ralys, R. V.; Zaitsau, D. H.; Stark, A. Biomass-derived platform chemicals: Thermodynamic studies on the conversion of 5-hydroxymethylfurfural into bulk intermediates. Ind. Eng. Chem. Res. 2009, 48, 10087−10093.

dehydration of glucose to 5-hydroxymethylfurfural on acidic mesoporous tantalum phosphate. Appl. Catal., B 2014, 144, 22−28. (32) Zhang, X. M.; Murria, P.; Jiang, Y.; Xiao, W. H.; Kenttämaa, H. I.; Abu-Omar, M. M.; Mosier, N. S. Maleic acid and aluminum chloride catalyzed conversion of glucose to 5-(hydroxymethyl) furfural and levulinic acid in aqueous media. Green Chem. 2016, 18, 5219−5229. (33) Tong, X. L.; Li, M. R.; Yan, N.; Ma, Y.; Dyson, P. J.; Li, Y. D. Defunctionalization of fructose and sucrose: Iron-catalyzed production of 5-hydroxymethylfurfural from fructose and sucrose. Catal. Today 2011, 175, 524−527. (34) Wang, C.; Fu, L. T.; Tong, X. L.; Yang, Q. W.; Zhang, W. Efficient and selective conversion of sucrose to 5-hydroxymethylfurfural promoted by ammonium halides under mild conditions. Carbohydr. Res. 2012, 347, 182−185. (35) Jadhav, A. H.; Kim, H.; Hwang, I. T. An efficient and heterogeneous recyclable silicotungstic acid with modified acid sites as a catalyst for conversion of fructose and sucrose into 5hydroxymethylfurfural in superheated water. Bioresour. Technol. 2013, 132, 342−350. (36) Pérez-Maqueda, J.; Arenas-Ligioiz, I.; López, Ó .; FernándezBolaños, J. G. Eco-friendly preparation of 5-hydroxymethylfurfural from sucrose using ion-exchange resins. Chem. Eng. Sci. 2014, 109, 244−250. (37) Kreissl, H. T.; Nakagawa, K.; Peng, Y. K.; Koito, Y.; Zheng, J. L.; Tsang, S. C. E. Niobium oxides: Correlation of acidity with structure and catalytic performance in sucrose conversion to 5-hydroxymethylfurfural. J. Catal. 2016, 338, 329−339. (38) Shi, X. L.; Zhang, M.; Lin, H.; Tao, M.; Li, Y.; Zhang, W. Bifunctional polyacrylonitrile fiber-mediated conversion of sucrose to 5-hydroxymethylfurfural in mixed-aqueous systems. Chem. - Asian J. 2015, 10, 752−758. (39) Beckerle, K.; Okuda, J. Conversion of glucose and cellobiose into 5-hydroxymethylfurfural (HMF) by rare earth metal salts in N,N′dimethylacetamide (DMA). J. Mol. Catal. A: Chem. 2012, 356, 158− 164. (40) Kimura, H.; Yoshida, K.; Uosaki, Y.; Nakahara, M. Effect of water content on conversion of cellobiose into 5-hydroxymethyl-2furaldehyde in a dimethyl sulfoxide-water mixture. J. Phys. Chem. A 2013, 117, 10987−10996. (41) Wang, J. J.; Ren, J. W.; Liu, X. H.; Xi, J. X.; Xia, Q. N.; Zu, Y. H.; Lu, G. Z.; Wang, Y. Q. Direct conversion of carbohydrates to 5hydroxymethylfurfural using Sn-Mont catalyst. Green Chem. 2012, 14, 2506−2512. (42) Hu, L.; Zhao, G.; Tang, X.; Wu, Z.; Xu, J. X.; Lin, L.; Liu, S. J. Catalytic conversion of carbohydrates into 5-hydroxymethylfurfural over cellulose-derived carbonaceous catalyst in ionic liquid. Bioresour. Technol. 2013, 148, 501−507. (43) Hu, L.; Wu, Z.; Xu, J. X.; Sun, Y.; Lin, L.; Liu, S. J. Zeolitepromoted transformation of glucose into 5-hydroxymethylfurfural in ionic liquid. Chem. Eng. J. 2014, 244, 137−144. (44) Hu, S. Q.; Zhang, Z. F.; Zhou, Y. X.; Song, J. L.; Fan, H. L.; Han, B. X. Direct conversion of inulin to 5-hydroxymethylfurfural in biorenewable ionic liquids. Green Chem. 2009, 11, 873−877. (45) Benoit, M.; Brissonnet, Y.; Guelou, E.; de Oliveira Vigier, K.; Barrault, J.; Jerome, F. Acid-catalyzed dehydration of fructose and inulin with glycerol or glycerol carbonate as renewably sourced cosolvent. ChemSusChem 2010, 3, 1304−1309. (46) Yi, Y. B.; Lee, J. W.; Hong, S. S.; Choi, Y. H.; Chung, C. H. Acid-mediated production of hydroxymethylfurfural from raw plant biomass with high inulin in an ionic liquid. J. Ind. Eng. Chem. 2011, 17, 6−9. (47) Shen, X.; Wang, Y. X.; Hu, C. W.; Qian, K.; Ji, Z.; Jin, M. Onepot conversion of inulin to furan derivatives catalyzed by sulfated TiO2/mordenite solid acid. ChemCatChem 2012, 4, 2013−2019. (48) Xie, H. B.; Zhao, Z. B.; Wang, Q. Catalytic conversion of inulin and fructose into 5-hydroxymethylfurfural by lignosulfonic acid in ionic liquids. ChemSusChem 2012, 5, 901−905. (49) De, S.; Dutta, S.; Saha, B. Microwave assisted conversion of carbohydrates and biopolymers to 5-hydroxymethylfurfural with 2973

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

hydrogenation with ruthenium supported on carbon. ChemSusChem 2013, 6, 1158−1162. (86) Chidambaram, M.; Bell, A. T. A two-step approach for the catalytic conversion of glucose to 2,5-dimethylfuran in ionic liquids. Green Chem. 2010, 12, 1253−1262. (87) Mani, C. M.; Braun, M.; Molinari, V.; Antonietti, M.; Fechler, N. A high-throughput composite catalyst based on nickel carbon cubes for the hydrogenation of 5-hydroxymethylfurfural to 2,5-dimethylfuran. ChemCatChem 2017, 9, 3388−3394. (88) Kong, X.; Zhu, Y. F.; Zheng, H. Y.; Dong, F.; Zhu, Y. L.; Li, Y. W. Switchable synthesis of 2,5-dimethylfuran and 2,5-dihydroxymethyltetrahydrofuran from 5-hydroxymethylfurfural over Raney Ni catalyst. RSC Adv. 2014, 4, 60467−60472. (89) Zhou, H. C.; Song, J. L.; Meng, Q. L.; He, Z. H.; Jiang, Z. W.; Zhou, B. W.; Liu, H. Z.; Han, B. X. Cooperative catalysis of Pt/C and acid resin for the production of 2,5-dimethyltetrahydrofuran from biomass derived 2,5-hexanedione under mild conditions. Green Chem. 2016, 18, 220−225. (90) Jackson, M. A.; Appell, M.; Blackburn, J. A. Hydrodeoxygenation of fructose to 2,5-dimethyltetrahydrofuran using a sulfur poisoned Pt/C catalyst. Ind. Eng. Chem. Res. 2015, 54, 7059−7066. (91) Grochowski, M. R.; Yang, W.; Sen, A. Mechanistic study of a one-step catalytic conversion of fructose to 2,5-dimethyltetrahydrofuran. Chem. - Eur. J. 2012, 18, 12363−12371. (92) Yang, W. R.; Sen, A. One-step catalytic transformation of carbohydrates and cellulosic biomass to 2,5-dimethyltetrahydrofuran for liquid fuels. ChemSusChem 2010, 3, 597−603. (93) Luijkx, G. C. A.; Huck, N. P. M.; van Rantwijk, F.; Maat, L.; van Bekkum, H. Ether formation in the hydrogenolysis of hydroxymethylfurfural over palladium catalysts in alcoholic solution. Heterocycles 2009, 77, 1037−1044. (94) Wu, W. P.; Xu, Y. J.; Zhu, R.; Cui, M. S.; Li, X. L.; Deng, J.; Fu, Y. Selective conversion of 5-hydroxymethylfuraldehyde using Cp*Ir catalysts in aqueous formate buffer solution. ChemSusChem 2016, 9, 1209−1215. (95) Xu, Y. J.; Shi, J.; Wu, W. P.; Zhu, R.; Li, X. L.; Deng, J.; Fu, Y. Effect of Cp*Iridium(III) complex and acid co-catalyst on conversion of furfural compounds to cyclopentanones or straight chain ketones. Appl. Catal., A 2017, 543, 266−273. (96) Xu, Z. W.; Yan, P. F.; Li, H. X.; Liu, K. R.; Liu, X. M.; Jia, S. Y.; Zhang, Z. C. Active Cp*Iridium(III) complex with ortho-hydroxyl group functionalized bipyridine ligand containing an electron-donating group for the production of diketone from 5-HMF. ACS Catal. 2016, 6, 3784−3788. (97) Gupta, K.; Tyagi, D.; Dwivedi, A. D.; Mobin, S. M.; Singh, S. K. Catalytic transformation of bio-derived furans to valuable ketoacids and diketones by water-soluble ruthenium catalysts. Green Chem. 2015, 17, 4618−4627. (98) Xu, Z. W.; Yan, P. F.; Xu, W. J.; Liu, X. M.; Xia, Z.; Chung, B.; Jia, S. Y.; Zhang, Z. C. Hydrogenation/hydrolytic ring opening of 5HMF by Cp*Iridium(III) half-sandwich complexes for bioketones synthesis. ACS Catal. 2015, 5, 788−792. (99) Liu, F.; Audemar, M.; de Oliveira Vigier, K.; Clacens, J. M.; de Campo, F.; Jerome, F. Palladium/carbon dioxide cooperative catalysis for the production of diketone derivatives from carbohydrates. ChemSusChem 2014, 7, 2089−2093. (100) Liu, L.; Ye, X. P.; Bozell, J. J. A comparative review of petroleum-based and bio-based acrolein production. ChemSusChem 2012, 5, 1162−1180. (101) Buntara, T.; Melián-Cabrera, I.; Tan, Q. H.; Fierro, J. L. G.; Neurock, M.; de Vries, J. G.; Heeres, H. J. Catalyst studies on the ring opening of tetrahydrofuran-dimethanol to 1,2,6-hexanetriol. Catal. Today 2013, 210, 106−116. (102) Nolan, M. R.; Sun, G.; Shanks, B. H. On the selective acidcatalysed dehydration of 1,2,6-hexanetriol. Catal. Sci. Technol. 2014, 4, 2260−2266. (103) Chen, K.; Koso, S.; Kubota, T.; Nakagawa, Y.; Tomishige, K. Chemoselective hydrogenolysis of tetrahydropyran-2-methanol to 1,6-

(68) Scholz, D.; Aellig, C.; Hermans, I. Catalytic transfer hydrogenation/hydrogenolysis for reductive upgrading of furfural and 5(hydroxymethyl)furfural. ChemSusChem 2014, 7, 268−275. (69) Chatterjee, M.; Ishizaka, T.; Kawanami, H. Hydrogenation of 5hydroxymethylfurfural in supercritical carbon dioxide/water: A tunable approach to dimethylfuran selectivity. Green Chem. 2014, 16, 1543− 1551. (70) Elangovan, S.; Topf, C.; Fischer, S.; Jiao, H.; Spannenberg, A.; Baumann, W.; Ludwig, R.; Junge, K.; Beller, M. Selective catalytic hydrogenations of nitriles, ketones, and aldehydes by well-defined manganese pincer complexes. J. Am. Chem. Soc. 2016, 138, 8809− 8814. (71) Perret, N.; Grigoropoulos, A.; Zanella, M.; Manning, T. D.; Claridge, J. B.; Rosseinsky, M. J. Catalytic response and stability of nickel/alumina for the hydrogenation of 5-hydroxymethylfurfural in water. ChemSusChem 2016, 9, 521−531. (72) Chen, J. Z.; Liu, R. L.; Guo, Y. Y.; Chen, L. M.; Gao, H. Selective hydrogenation of biomass based 5-hydroxymethylfurfural over catalyst of palladium immobilized on amine-functionalized metalorganic frameworks. ACS Catal. 2015, 5, 722−733. (73) Yang, Y. L.; Du, Z. T.; Ma, J. P.; Lu, F.; Zhang, J. J.; Xu, J. Biphasic catalytic conversion of fructose by continuous hydrogenation of HMF over a hydrophobic ruthenium catalyst. ChemSusChem 2014, 7, 1352−1356. (74) Nakagawa, Y.; Takada, K.; Tamura, M.; Tomishige, K. Total hydrogenation of furfural and 5-hydroxymethylfurfural over supported Pd-Ir alloy catalyst. ACS Catal. 2014, 4, 2718−2726. (75) Nakagawa, Y.; Tomishige, K. Total hydrogenation of furan derivatives over silica-supported Ni-Pd alloy catalyst. Catal. Commun. 2010, 12, 154−156. (76) Connolly, T. J.; Considine, J. L.; Ding, Z. X.; Forsatz, B.; Jennings, M. N.; MacEwan, M. F.; McCoy, K. M.; Place, D. W.; Sharma, A.; Sutherland, K. Efficient synthesis of 8-oxa-3-azabicyclo[3.2.1]octane hydrochloride. Org. Process Res. Dev. 2010, 14, 459−465. (77) Yang, P. P.; Xia, Q. N.; Liu, X. H.; Wang, Y. Q. Catalytic transfer hydrogenation/hydrogenolysis of 5-hydroxymethylfurfural to 2,5dimethylfuran over Ni-Co/C catalyst. Fuel 2017, 187, 159−166. (78) Luo, J.; Monai, M.; Wang, C.; Lee, J. D.; Duchoň, T.; Dvořaḱ , F.; Matolín, V.; Murray, C. B.; Fornasiero, P.; Gorte, R. J. Unraveling the surface state and composition of highly selective nanocrystalline Ni−Cu alloy catalysts for hydrodeoxygenation of HMF. Catal. Sci. Technol. 2017, 7, 1735−1743. (79) Li, H.; Zhao, W. F.; Riisager, A.; Saravanamurugan, S.; Wang, Z. W.; Fang, Z.; Yang, S. A Pd-Catalyzed in situ domino process for mild and quantitative production of 2,5-dimethylfuran directly from carbohydrates. Green Chem. 2017, 19, 2101−2106. (80) Guo, W. W.; Liu, H. Y.; Zhang, S. Q.; Han, H. L.; Liu, H. Z.; Jiang, T.; Han, B. X.; Wu, T. B. Efficient hydrogenolysis of 5hydroxymethylfurfural to 2,5-dimethylfuran over a cobalt and copper bimetallic catalyst on N-graphene-modified Al2O3. Green Chem. 2016, 18, 6222−6228. (81) Chen, M. Y.; Chen, C. B.; Zada, B.; Fu, Y. Perovskite type oxide supported Ni ctalysts for the production of 2,5-dimethylfuran from biomass-derived 5-hydroxymethylfurfural. Green Chem. 2016, 18, 3858−3866. (82) Kong, X.; Zhu, Y. F.; Zheng, H. Y.; Li, X. Q.; Zhu, Y. L.; Li, Y. W. Ni nanoparticles inlaid nickel phyllosilicate as a metal-acid bifunctional catalyst for low-temperature hydrogenolysis reactions. ACS Catal. 2015, 5, 5914−5920. (83) Kong, X.; Zheng, R. X.; Zhu, Y. F.; Ding, G. Q.; Zhu, Y. L.; Li, Y. W. Rational design of Ni-based catalysts derived from hydrotalcite for selective hydrogenation of 5-hydroxymethylfurfural. Green Chem. 2015, 17, 2504−2514. (84) Huang, Y. B.; Chen, M. Y.; Yan, L.; Guo, Q. X.; Fu, Y. Nickeltungsten carbide catalysts for the production of 2,5-dimethylfuran from biomass-derived molecules. ChemSusChem 2014, 7, 1068−1072. (85) Jae, J.; Zheng, W.; Lobo, R. F.; Vlachos, D. G. Production of dimethylfuran from hydroxymethylfurfural through catalytic transfer 2974

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis hexanediol over rhenium-modified carbon-supported rhodium catalysts. ChemCatChem 2010, 2, 547−555. (104) Buntara, T.; Noel, S.; Phua, P. H.; Melián-Cabrera, I.; de Vries, J. G.; Heeres, H. J. From 5-hydroxymethylfurfural (HMF) to polymer precursors: Catalyst screening studies on the conversion of 1,2,6hexanetriol to 1,6-hexanediol. Top. Catal. 2012, 55, 612−619. (105) Xiao, B.; Zheng, M. Y.; Li, X. S.; Pang, J. F.; Sun, R. Y.; Wang, H.; Pang, X. L.; Wang, A. Q.; Wang, X. D.; Zhang, T. Synthesis of 1,6hexanediol from HMF over double-layered catalysts of Pd/SiO2+IrReOx/SiO2 in a fixed-bed reactor. Green Chem. 2016, 18, 2175−2184. (106) Tuteja, J.; Choudhary, H.; Nishimura, S.; Ebitani, K. Direct synthesis of 1,6-hexanediol from HMF over a heterogeneous Pd/ZrP catalyst using formic acid as hydrogen source. ChemSusChem 2014, 7, 96−100. (107) Buntara, T.; Noel, S.; Phua, P. H.; Melián-Cabrera, I.; de Vries, J. G.; Heeres, H. J. Caprolactam from renewable resources: Catalytic conversion of 5-hydroxymethylfurfural into caprolactone. Angew. Chem., Int. Ed. 2011, 50, 7083−7087. (108) Ohyama, J.; Ohira, Y.; Satsuma, A. Hydrogenative ringrearrangement of biomass derived 5-(hydroxymethyl)furfural to 3(hydroxymethyl)cyclopentanol using combination catalyst systems of Pt/SiO2 and lanthanoid oxides. Catal. Sci. Technol. 2017, 7, 2947− 2953. (109) Ohyama, J.; Kanao, R.; Esaki, A.; Satsuma, A. Conversion of 5hydroxymethylfurfural to a cyclopentanone derivative by ring rearrangement over supported Au nanoparticles. Chem. Commun. 2014, 50, 5633−5636. (110) Ohyama, J.; Kanao, R.; Ohira, Y.; Satsuma, A. The effect of heterogeneous acid-base catalysis on conversion of 5-hydroxymethylfurfural into a cyclopentanone derivative. Green Chem. 2016, 18, 676− 680. (111) Timko, J. M.; Cram, D. J. Furanyl unit in host compounds. J. Am. Chem. Soc. 1974, 96, 7159−7160. (112) Balakrishnan, M.; Sacia, E. R.; Bell, A. T. Etherification and reductive etherification of 5-(hydroxymethyl)furfural: 5(alkoxymethyl)furfurals and 2,5-bis(alkoxymethyl)furans as potential bio-diesel candidates. Green Chem. 2012, 14, 1626−1634. (113) Cao, Q.; Liang, W. Y.; Guan, J.; Wang, L.; Qu, Q.; Zhang, X. Z.; Wang, X. C.; Mu, X. D. Catalytic synthesis of 2,5-bismethoxymethylfuran: A promising cetane number improver for diesel. Appl. Catal., A 2014, 481, 49−53. (114) Gelmini, A.; Albonetti, S.; Cavani, F.; Cesari, C.; Lolli, A.; Zanotti, V.; Mazzoni, R. Oxidant free one-pot transformation of biobased 2,5-bis-hydroxymethylfuran into α-6-hydroxy-6-methyl-4-enyl2H-pyran-3-one in water. Appl. Catal., B 2016, 180, 38−43. (115) Lecomte, J.; Finiels, A.; Geneste, P.; Moreau, C. Selective hydroxymethylation of furfuryl alcohol with aqueous formaldehyde in the presence of dealuminated mordenites. Appl. Catal., A 1998, 168, 235−241. (116) Lecomte, J.; Finiels, A.; Geneste, P.; Moreau, C. Kinetics of furfuryl alcohol hydroxymethylation with aqueous formaldehyde over a highly dealuminated H-mordenite. J. Mol. Catal. A: Chem. 1998, 133, 283−288. (117) Gandini, A. The irruption of polymers from renewable resources on the scene of macromolecular science and technology. Green Chem. 2011, 13, 1061−1083. (118) Jang, N. R.; Kim, H. R.; Hou, C. T.; Kim, B. S. Novel biobased photo-crosslinked polymer networks prepared from vegetable oil and 2,5-furan diacrylate. Polym. Adv. Technol. 2013, 24, 814−818. (119) Moreau, C.; Belgacem, M. N.; Gandini, A. Recent catalytic advances in the chemistry of substituted furans from carbohydrates and in the ensuing polymers. Top. Catal. 2004, 27, 11−30. (120) Delidovich, I.; Hausoul, P. J.; Deng, L.; Pfutzenreuter, R.; Rose, M.; Palkovits, R. Alternative monomers based on lignocellulose and their use for polymer production. Chem. Rev. 2016, 116, 1540−1599. (121) Nakagawa, Y.; Tamura, M.; Tomishige, K. Catalytic reduction of biomass-derived furanic compounds with hydrogen. ACS Catal. 2013, 3, 2655−2668.

(122) Hu, L.; Tang, X.; Xu, J. X.; Wu, Z.; Lin, L.; Liu, S. J. Selective transformation of 5-hydroxymethylfurfural into the liquid fuel 2,5dimethylfuran over carbon-supported ruthenium. Ind. Eng. Chem. Res. 2014, 53, 3056−3064. (123) Hu, L.; Lin, L.; Liu, S. J. Chemoselective hydrogenation of biomass-derived 5-hydroxymethylfurfural into the liquid biofuel 2,5dimethylfuran. Ind. Eng. Chem. Res. 2014, 53, 9969−9978. (124) De, S.; Saha, B.; Luque, R. Hydrodeoxygenation processes: Advances on catalytic transformations of biomass-derived platform chemicals into hydrocarbon fuels. Bioresour. Technol. 2015, 178, 108− 118. (125) Jain, A. B.; Vaidya, P. D. Kinetics of catalytic hydrogenation of 5-hydroxymethylfurfural to 2,5-bis-hydroxymethylfuran in aqueous solution over Ru/C. Int. J. Chem. Kinet. 2016, 48, 318−328. (126) Mäki-Arvela, P.; Hájek, J.; Salmi, T.; Murzin, D. Y. Chemoselective hydrogenation of carbonyl compounds over heterogeneous catalysts. Appl. Catal., A 2005, 292, 1−49. (127) Ide, M. S.; Hao, B.; Neurock, M.; Davis, R. J. Mechanistic insights on the hydrogenation of α,β-unsaturated ketones and aldehydes to unsaturated alcohols over metal catalysts. ACS Catal. 2012, 2, 671−683. (128) Claus, P. Selective hydrogenation of α,β-unsaturated aldehydes and other CO and CC bonds containing compounds. Top. Catal. 1998, 5, 51−62. (129) Ponec, V. On the role of promoters in hydrogenations on metals: α,β-Unsaturated aldehydes and ketones. Appl. Catal., A 1997, 149, 27−48. (130) Yuan, Y.; Yao, S. Y.; Wang, M. N.; Lou, S. J.; Yan, N. Recent progress in chemoselective hydrogenation of α,β-unsaturated aldehyde to unsaturated alcohol over nanomaterials. Curr. Org. Chem. 2013, 17, 400−413. (131) Milone, C. Selective hydrogenation of α,β-unsaturated ketones to α,β-unsaturated alcohols on gold-supported catalysts. J. Catal. 2004, 222, 348−356. (132) Gilkey, M. J.; Xu, B. J. Heterogeneous catalytic transfer hydrogenation as an effective pathway in biomass upgrading. ACS Catal. 2016, 6, 1420−1436. (133) Lichtenthaler, F. W.; Brust, A.; Cuny, E. Sugar-derived building blocks Part 26: Hydrophilic pyrroles, pyridazines and diazepinones from D-fructose and isomaltulose. Green Chem. 2001, 3, 201−209. (134) Goswami, S.; Dey, S.; Jana, S. Design and synthesis of a unique ditopic macrocyclic fluorescent receptor containing furan ring as a spacer for the recognition of dicarboxylic acids. Tetrahedron 2008, 64, 6358−6363. (135) Saha, B.; Bohn, C. M.; Abu-Omar, M. M. Zinc-assisted hydrodeoxygenation of biomass-derived 5-hydroxymethylfurfural to 2,5-dimethylfuran. ChemSusChem 2014, 7, 3095−3101. (136) Lin, R. C.; Cheng, J.; Ding, L. K.; Song, W. L.; Zhou, J. H.; Cen, K. F. Sodium borohydride removes aldehyde inhibitors for enhancing biohydrogen fermentation. Bioresour. Technol. 2015, 197, 323−328. (137) Aellig, C.; Jenny, F.; Scholz, D.; Wolf, P.; Giovinazzo, I.; Kollhoff, F.; Hermans, I. Combined 1,4-butanediol lactonization and transfer hydrogenation/hydrogenolysis of furfural-derivatives under continuous flow conditions. Catal. Sci. Technol. 2014, 4, 2326−2331. (138) Pasini, T.; Solinas, G.; Zanotti, V.; Albonetti, S.; Cavani, F.; Vaccari, A.; Mazzanti, A.; Ranieri, S.; Mazzoni, R. Substrate and product role in the Shvo’s catalyzed selective hydrogenation of the platform bio-based chemical 5-hydroxymethylfurfural. Dalton Trans. 2014, 43, 10224−10234. (139) Op de Beeck, B.; Dusselier, M.; Geboers, J.; Holsbeek, J.; Morré, E.; Oswald, S.; Giebeler, L.; Sels, B. F. Direct catalytic conversion of cellulose to liquid straight-chain alkanes. Energy Environ. Sci. 2015, 8, 230−240. (140) Han, J. S.; Kim, Y. H.; Jang, H. S.; Hwang, S. Y.; Jegal, J.; Kim, J. W.; Lee, Y. S. Heterogeneous zirconia-supported ruthenium catalyst for highly selective hydrogenation of 5-hydroxymethyl-2-furaldehyde to 2,5-bis(hydroxymethyl)furans in various n-alcohol solvents. RSC Adv. 2016, 6, 93394−93397. 2975

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis

(158) Tamura, M.; Tokonami, K.; Nakagawa, Y.; Tomishige, K. Rapid synthesis of unsaturated alcohols under mild conditions by highly selective hydrogenation. Chem. Commun. 2013, 49, 7034−7036. (159) Shi, J. J.; Zhang, M. Y.; Du, W. C.; Ning, W. S.; Hou, Z. Y. SnO2-isolated Pt3Sn alloy on reduced graphene oxide: An efficient catalyst for selective hydrogenation of CO in unsaturated aldehydes. Catal. Sci. Technol. 2015, 5, 3108−3112. (160) Noyori, R.; Hashiguchi, S. Asymmetric transfer hydrogenation catalyzed by chiral ruthenium complexes. Acc. Chem. Res. 1997, 30, 97−102. (161) Ikariya, T.; Blacker, A. J. Asymmetric transfer hydrogenation of ketones with bifunctional transition metal-based molecular catalysts. Acc. Chem. Res. 2007, 40, 1300−1308. (162) Wang, D.; Astruc, D. The golden age of transfer hydrogenation. Chem. Rev. 2015, 115, 6621−6686. (163) Osatiashtiani, A.; Lee, A. F.; Wilson, K. Recent advances in the production of γ-valerolactone from biomass-derived feedstocks via heterogeneous catalytic transfer hydrogenation. J. Chem. Technol. Biotechnol. 2017, 92, 1125−1135. (164) Xue, Z. M.; Jiang, J. Y.; Li, G. F.; Zhao, W. C.; Wang, J. F.; Mu, T. C. Zirconium-cyanuric acid coordination polymer: Highly efficient catalyst for conversion of levulinic acid to γ-valerolactone. Catal. Sci. Technol. 2016, 6, 5374−5379. (165) Stolle, A.; Gallert, T.; Schmoger, C.; Ondruschka, B. Hydrogenation of citral: A wide-spread model reaction for selective reduction of α,β-unsaturated aldehydes. RSC Adv. 2013, 3, 2112− 2153. (166) Aramendía, M. A.; Borau, V.; Jiménez, C.; Marinas, J. M.; Ruiz, J. R.; Urbano, F. J. Liquid-phase heterogeneous catalytic transfer hydrogenation of citral on basic catalysts. J. Mol. Catal. A: Chem. 2001, 171, 153−158. (167) Aramendía, M. A.; Borau, V.; Jiménez, C.; Marinas, J. M.; Ruiz, J. R.; Urbano, F. J. Catalytic transfer hydrogenation of citral on calcined layered double hydroxides. Appl. Catal., A 2001, 206, 95−101. (168) Su, F. Z.; He, L.; Ni, J.; Cao, Y.; He, H. Y.; Fan, K. N. Efficient and chemoselective reduction of carbonyl compounds with supported gold catalysts under transfer hydrogenation conditions. Chem. Commun. 2008, 3531−3533. (169) Yuan, Q. Q.; Zhang, D. M.; van Haandel, L.; Ye, F. Y.; Xue, T.; Hensen, E. J. M.; Guan, Y. J. Selective liquid phase hydrogenation of furfural to furfuryl alcohol by Ru/Zr-MOFs. J. Mol. Catal. A: Chem. 2015, 406, 58−64. (170) Mariscal, R.; Maireles-Torres, P.; Ojeda, M.; Sádaba, I.; López Granados, M. Furfural: A renewable and versatile platform molecule for the synthesis of chemicals and fuels. Energy Environ. Sci. 2016, 9, 1144−1189. (171) Gao, Z.; Yang, L.; Fan, G. L.; Li, F. Promotional role of surface defects on carbon-supported Ru-based catalysts in transfer hydrogenation of furfural. ChemCatChem 2016, 8, 3769−3779. (172) Panagiotopoulou, P.; Martin, N.; Vlachos, D. G. Liquid-phase catalytic transfer hydrogenation of furfural over homogeneous Lewis acid-Ru/C catalysts. ChemSusChem 2015, 8, 2046−2054. (173) Villaverde, M. M.; Garetto, T. F.; Marchi, A. J. Liquid-phase transfer hydrogenation of furfural to furfuryl alcohol on Cu-Mg-Al catalysts. Catal. Commun. 2015, 58, 6−10. (174) Zhang, B.; Tang, M. H.; Yuan, J.; Wu, L. Support effect in Meerwein−Ponndorf−Verley reduction of benzaldehyde over supported zirconia catalysts. Chin. J. Catal. 2012, 33, 914−922. (175) Mollica, A.; Genovese, S.; Pinnen, F.; Stefanucci, A.; Curini, M.; Epifano, F. Ytterbium triflate catalysed Meerwein−Ponndorf− Verley (MPV) reduction. Tetrahedron Lett. 2012, 53, 890−892. (176) Jiménez-Sanchidrián, C.; Ruiz, J. R. Tin-containing hydrotalcite-like compounds as catalysts for the Meerwein−Ponndorf− Verley reaction. Appl. Catal., A 2014, 469, 367−372. (177) Bartley, J. K.; Xu, C.; Lloyd, R.; Enache, D. I.; Knight, D. W.; Hutchings, G. J. Simple method to synthesize high surface area magnesium oxide and its use as a heterogeneous base catalyst. Appl. Catal., B 2012, 128, 31−38.

(141) Alamillo, R.; Tucker, M.; Chia, M.; Pagán-Torres, Y.; Dumesic, J. A. The selective hydrogenation of biomass-derived 5-hydroxymethylfurfural using heterogeneous catalysts. Green Chem. 2012, 14, 1413−1419. (142) Chen, J. Z.; Lu, F.; Zhang, J. J.; Yu, W. Q.; Wang, F.; Gao, J.; Xu, J. Immobilized Ru clusters in nanosized mesoporous zirconium silica for the aqueous hydrogenation of furan derivatives at room temperature. ChemCatChem 2013, 5, 2822−2826. (143) Schiavo, V.; Descotes, G.; Mentech, J. Catalytic hydrogenation of 5-hydroxymethylfurfural in aqueous medium. Bull. Soc. Chim. Fr. 1991, 128, 704−711. (144) Chatterjee, M.; Ishizaka, T.; Kawanami, H. Selective hydrogenation of 5-hydroxymethylfurfural to 2,5-bis-(hydroxymethyl)furan using Pt/MCM-41 in an aqueous medium: A simple approach. Green Chem. 2014, 16, 4734−4739. (145) Liu, F.; Audemar, M.; de Oliveira Vigier, K.; Clacens, J. M.; de Campo, F.; Jérôme, F. Combination of Pd/C and Amberlyst-15 in a single reactor for the acid/hydrogenating catalytic conversion of carbohydrates to 5-hydroxy-2,5-hexanedione. Green Chem. 2014, 16, 4110−4114. (146) Cai, H. L.; Li, C. Z.; Wang, A. Q.; Zhang, T. Biomass into chemicals: One-pot production of furan-based diols from carbohydrates via tandem reactions. Catal. Today 2014, 234, 59−65. (147) Ohyama, J.; Hayashi, Y.; Ueda, K.; Yamamoto, Y.; Arai, S.; Satsuma, A. Effect of FeOx modification of Al2O3 on its supported Au catalyst for hydrogenation of 5-hydroxymethylfurfural. J. Phys. Chem. C 2016, 120, 15129−15136. (148) Ohyama, J.; Esaki, A.; Yamamoto, Y.; Arai, S.; Satsuma, A. Selective hydrogenation of 2-hydroxymethyl-5-furfural to 2,5-bis(hydroxymethyl)furan over gold sub-nano clusters. RSC Adv. 2013, 3, 1033−1036. (149) Kumalaputri, A. J.; Bottari, G.; Erne, P. M.; Heeres, H. J.; Barta, K. Tunable and selective conversion of 5-HMF to 2,5-furandimethanol and 2,5-dimethylfuran over copper-doped porous metal oxides. ChemSusChem 2014, 7, 2266−2275. (150) Zhu, Y. F.; Kong, X.; Zheng, H. Y.; Ding, G. Q.; Zhu, Y. L.; Li, Y. W. Efficient synthesis of 2,5-dihydroxymethylfuran and 2,5dimethylfuran from 5-hydroxymethylfurfural using mineral-derived Cu catalysts as versatile catalysts. Catal. Sci. Technol. 2015, 5, 4208− 4217. (151) Lima, S.; Chadwick, D.; Hellgardt, K. Towards sustainable hydrogenation of 5-(hydroxymethyl)furfural: A two-stage continuous process in aqueous media over RANEY® catalysts. RSC Adv. 2017, 7, 31401−31407. (152) Chen, B. F.; Li, F. B.; Huang, Z. J.; Yuan, G. Q. Carbon-coated Cu-Co bimetallic nanoparticles as selective and recyclable catalysts for production of biofuel 2,5-dimethylfuran. Appl. Catal., B 2017, 200, 192−199. (153) Yu, L. L.; He, L.; Chen, J.; Zheng, J. W.; Ye, L. M.; Lin, H. Q.; Yuan, Y. Z. Robust and recyclable nonprecious bimetallic nanoparticles on carbon nanotubes for the hydrogenation and hydrogenolysis of 5hydroxymethylfurfural. ChemCatChem 2015, 7, 1701−1707. (154) Srivastava, S.; Jadeja, G. C.; Parikh, J. Synergism studies on alumina-supported copper-nickel catalysts towards furfural and 5hydroxymethylfurfural hydrogenation. J. Mol. Catal. A: Chem. 2017, 426, 244−256. (155) Bottari, G.; Kumalaputri, A. J.; Krawczyk, K. K.; Feringa, B. L.; Heeres, H. J.; Barta, K. Copper-zinc alloy nanopowder: A robust precious-metal-free catalyst for the conversion of 5-hydroxymethylfurfural. ChemSusChem 2015, 8, 1323−1327. (156) Yao, S. X.; Wang, X. C.; Jiang, Y. J.; Wu, F.; Chen, X. G.; Mu, X. D. One-step conversion of biomass-derived 5-hydroxymethylfurfural to 1,2,6-hexanetriol over Ni-Co-Al mixed oxide catalysts under mild conditions. ACS Sustainable Chem. Eng. 2014, 2, 173−180. (157) Wang, G. H.; Hilgert, J.; Richter, F. H.; Wang, F.; Bongard, H. J.; Spliethoff, B.; Weidenthaler, C.; Schuth, F. Platinum-cobalt bimetallic nanoparticles in hollow carbon nanospheres for hydrogenolysis of 5-hydroxymethylfurfural. Nat. Mater. 2014, 13, 293−300. 2976

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis (178) Polshettiwar, V.; Varma, R. S. Revisiting the Meerwein− Ponndorf−Verley reduction: A sustainable protocol for transfer hydrogenation of aldehydes and ketones. Green Chem. 2009, 11, 1313−1316. (179) Miñambres, J. F.; Marinas, A.; Marinas, J. M.; Urbano, F. J. Activity and deactivation of catalysts based on zirconium oxide modified with metal chlorides in the MPV reduction of crotonaldehyde. Appl. Catal., B 2013, 140−141, 386−395. (180) Miñambres, J. F.; Marinas, A.; Marinas, J. M.; Urbano, F. J. Chemoselective crotonaldehyde hydrogen transfer reduction over pure and supported metal nitrates. J. Catal. 2012, 295, 242−253. (181) Axpuac, S.; Aramendía, M. A.; Hidalgo-Carrillo, J.; Marinas, A.; Marinas, J. M.; Montes-Jiménez, V.; Urbano, F. J.; Borau, V. Study of structure-performance relationships in Meerwein−Ponndorf−Verley reduction of crotonaldehyde on several magnesium and zirconiumbased systems. Catal. Today 2012, 187, 183−190. (182) Noller, H.; Lin, W. M. Activity and selectivity of Ni-Cu/Al2O3 catalysts for hydrogenation of crotonaldehyde and mechanism of hydrogenation. J. Catal. 1984, 85, 25−30. (183) Bogár, K.; Krumlinde, P.; Bacsik, Z.; Hedin, N.; Bäckvall, J. E. Heterogenized Wilkinson-type catalyst for transfer hydrogenation of carbonyl compounds. Eur. J. Org. Chem. 2011, 2011, 4409−4414. (184) Ishikawa, S.; Jones, D. R.; Iqbal, S.; Reece, C.; Morgan, D. J.; Willock, D. J.; Miedziak, P. J.; Bartley, J. K.; Edwards, J. K.; Murayama, T.; Ueda, W.; Hutchings, G. J. Identification of the catalytically active component of Cu-Zr-O catalyst for the hydrogenation of levulinic acid to γ-valerolactone. Green Chem. 2017, 19, 225−236. (185) Al-Shaal, M. G.; Calin, M.; Delidovich, I.; Palkovits, R. Microwave-assisted reduction of levulinic acid with alcohols producing γ-valerolactone in the presence of a Ru/C catalyst. Catal. Commun. 2016, 75, 65−68. (186) Lv, J. K.; Rong, Z. M.; Wang, Y.; Xiu, J. H.; Wang, Y.; Qu, J. P. Highly efficient conversion of biomass-derived levulinic acid into γvalerolactone over Ni/MgO catalyst. RSC Adv. 2015, 5, 72037−72045. (187) Chia, M.; Dumesic, J. A. Liquid-phase catalytic transfer hydrogenation and cyclization of levulinic acid and its esters to γvalerolactone over metal oxide catalysts. Chem. Commun. 2011, 47, 12233−12235. (188) Kopetzki, D.; Antonietti, M. Transfer hydrogenation of levulinic acid under hydrothermal conditions catalyzed by sulfate as a temperature-switchable base. Green Chem. 2010, 12, 656−660. (189) Zheng, J.; Zhu, J.; Xu, X.; Wang, W.; Li, J.; Zhao, Y.; Tang, K.; Song, Q.; Qi, X.; Kong, D.; Tang, Y. Continuous hydrogenation of ethyl levulinate to γ-valerolactone and 2-methyl tetrahydrofuran over alumina doped Cu/SiO2 catalyst: the potential of commercialization. Sci. Rep. 2016, 6, 28898. (190) Valekar, A. H.; Cho, K. H.; Chitale, S. K.; Hong, D. Y.; Cha, G. Y.; Lee, U. H.; Hwang, D. W.; Serre, C.; Chang, J. S.; Hwang, Y. K. Catalytic transfer hydrogenation of ethyl levulinate to γ-valerolactone over zirconium-based metal-organic frameworks. Green Chem. 2016, 18, 4542−4552. (191) He, J.; Li, H.; Lu, Y. M.; Liu, Y. X.; Wu, Z. B.; Hu, D. Y.; Yang, S. Cascade catalytic transfer hydrogenation-cyclization of ethyl levulinate to γ-valerolactone with Al-Zr mixed oxides. Appl. Catal., A 2016, 510, 11−19. (192) He, J.; Li, H.; Liu, Y. X.; Zhao, W. F.; Yang, T. T.; Xue, W.; Yang, S. Catalytic transfer hydrogenation of ethyl levulinate into γvalerolactone over mesoporous Zr/B mixed oxides. J. Ind. Eng. Chem. 2016, 43, 133−141. (193) Song, J. L.; Wu, L. Q.; Zhou, B. W.; Zhou, H. C.; Fan, H. L.; Yang, Y. Y.; Meng, Q. L.; Han, B. X. New porous Zr-containing catalyst with phenate group: Efficient catalyst for catalytic transfer hydrogenation of ethyl levulinate to γ-valerolactone. Green Chem. 2015, 17, 1626−1632. (194) Grasemann, M.; Laurenczy, G. Formic acid as a hydrogen source: Recent developments and future trends. Energy Environ. Sci. 2012, 5, 8171−8181.

(195) Guzmán, I.; Heras, A.; Güemez, M. B.; Iriondo, A.; Cambra, J. F.; Requies, J. Levulinic acid production using solid-acid catalysis. Ind. Eng. Chem. Res. 2016, 55, 5139−5144. (196) Morone, A.; Apte, M.; Pandey, R. A. Levulinic acid production from renewable waste resources: Bottlenecks, potential remedies, advancements and applications. Renewable Sustainable Energy Rev. 2015, 51, 548−565. (197) Li, K. X.; Bai, L. L.; Amaniampong, P. N.; Jia, X. L.; Lee, J. M.; Yang, Y. H. One-pot transformation of cellobiose to formic acid and levulinic acid over ionic-liquid-based polyoxometalate hybrids. ChemSusChem 2014, 7, 2670−2677. (198) Weingarten, R.; Kim, Y. T.; Tompsett, G. A.; Fernandez, A.; Han, K. S.; Hagaman, E. W.; Conner, W. C.; Dumesic, J. A.; Huber, G. W. Conversion of glucose into levulinic acid with solid metal(IV) phosphate catalysts. J. Catal. 2013, 304, 123−134. (199) Thananatthanachon, T.; Rauchfuss, T. B. Efficient route to hydroxymethylfurans from sugars via transfer hydrogenation. ChemSusChem 2010, 3, 1139−1141. (200) Thananatthanachon, T.; Rauchfuss, T. B. Efficient production of the liquid fuel 2,5-dimethylfuran from fructose using formic acid as a reagent. Angew. Chem., Int. Ed. 2010, 49, 6616−6618. (201) Jae, J.; Zheng, W. Q.; Karim, A. M.; Guo, W.; Lobo, R. F.; Vlachos, D. G. The role of Ru and RuO2 in the catalytic transfer hydrogenation of 5-hydroxymethylfurfural for the production of 2,5dimethylfuran. ChemCatChem 2014, 6, 848−856. (202) Nagpure, A. S.; Venugopal, A. K.; Lucas, N.; Manikandan, M.; Thirumalaiswamy, R.; Chilukuri, S. Renewable fuels from biomassderived compounds: Ru-containing hydrotalcites as catalysts for conversion of HMF to 2,5-dimethylfuran. Catal. Sci. Technol. 2015, 5, 1463−1472. (203) Hao, W. W.; Li, W. F.; Tang, X.; Zeng, X. H.; Sun, Y.; Liu, S. J.; Lin, L. Catalytic transfer hydrogenation of biomass-derived 5hydroxymethyl furfural to the building block 2,5-bishydroxymethyl furan. Green Chem. 2016, 18, 1080−1088. (204) Li, H.; He, J.; Riisager, A.; Saravanamurugan, S.; Song, B.; Yang, S. Acid-base bifunctional zirconium N-alkyltriphosphate nanohybrid for hydrogen transfer of biomass-derived carboxides. ACS Catal. 2016, 6, 7722−7727. (205) Li, H.; Fang, Z.; He, J.; Yang, S. Orderly layered Zrbenzylphosphonate nanohybrids for efficient acid-base-mediated bifunctional/cascade catalysis. ChemSusChem 2017, 10, 681−686. (206) Li, H.; Liu, X. F.; Yang, T. T.; Zhao, W. F.; Saravanamurugan, S.; Yang, S. Porous zirconium-furandicarboxylate microspheres for efficient redox conversion of biofuranics. ChemSusChem 2017, 10, 1761−1770. (207) Wang, F.; Zhang, Z. H. Catalytic transfer hydrogenation of furfural into furfuryl alcohol over magnetic γ-Fe2O3@HAP catalyst. ACS Sustainable Chem. Eng. 2017, 5, 942−947. (208) Wang, G.-H.; Deng, X.; Gu, D.; Chen, K.; Tüysüz, H.; Spliethoff, B.; Bongard, H.-J.; Weidenthaler, C.; Schmidt, W.; Schüth, F. Co3O4 nanoparticles supported on mesoporous carbon for selective transfer hydrogenation of α,β-unsaturated aldehydes. Angew. Chem., Int. Ed. 2016, 55, 11101−11105. (209) Hao, C. H.; Guo, X. N.; Pan, Y. T.; Chen, S.; Jiao, Z. F.; Yang, H.; Guo, X. Y. Visible-light-driven selective photocatalytic hydrogenation of cinnamaldehyde over Au/SiC catalysts. J. Am. Chem. Soc. 2016, 138, 9361−9364. (210) Wang, T.; Zhang, J. H.; Xie, W. X.; Tang, Y. J.; Guo, D. L.; Ni, Y. H. Catalytic transfer hydrogenation of biobased HMF to 2,5-bis(hydroxymethyl)furan over Ru/Co3O4. Catalysts 2017, 7, 92−99. (211) Gao, Z.; Fan, G. L.; Yang, L.; Li, F. Double-active sites cooperatively catalyzed transfer hydrogenation of ethyl levulinate over a ruthenium-based catalyst. Mol. Catal. 2017, 442, 181−190. (212) de Graauw, C. F.; Peters, J. A.; van Bekkum, H.; Huskens, J. Meerwein−Ponndorf−Verley reductions and oppenauer oxidations: An integrated approach. Synthesis 1994, 1994, 1007−1017. (213) van der Waal, J. C.; Kunkeler, P. J.; Tan, K.; van Bekkum, H. Zeolite titanium beta: A selective catalyst for the gas-phase Meerwein− 2977

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis Ponndorf−Verley, and Oppenauer reactions. J. Catal. 1998, 173, 74− 83. (214) Tang, X.; Wei, J. N.; Ding, N.; Sun, Y.; Zeng, X. H.; Hu, L.; Liu, S. J.; Lei, T. Z.; Lin, L. Chemoselective hydrogenation of biomass derived 5-hydroxymethylfurfural to diols: Key intermediates for sustainable chemicals, materials and fuels. Renewable Sustainable Energy Rev. 2017, 77, 287−296. (215) Hu, L.; Yang, M.; Xu, N.; Xu, J. X.; Zhou, S. Y.; Chu, X. Z.; Zhao, Y. J. Selective transformation of biomass-derived 5-hydroxymethylfurfural into 2,5-dihydroxymethylfuran via catalytic transfer hydrogenation over magnetic zirconium hydroxides. Korean J. Chem. Eng. 2018, 35, 99−109. (216) Pasini, T.; Lolli, A.; Albonetti, S.; Cavani, F.; Mella, M. Methanol as a clean and efficient H-transfer reactant for carbonyl reduction: Scope, limitations, and reaction mechanism. J. Catal. 2014, 317, 206−219. (217) Yim, H.; Haselbeck, R.; Niu, W.; Pujol-Baxley, C.; Burgard, A.; Boldt, J.; Khandurina, J.; Trawick, J. D.; Osterhout, R. E.; Stephen, R.; Estadilla, J.; Teisan, S.; Schreyer, H. B.; Andrae, S.; Yang, T. H.; Lee, S. Y.; Burk, M. J.; van Dien, S. Metabolic engineering of Escherichia coli for direct production of 1,4-butanediol. Nat. Chem. Biol. 2011, 7, 445− 452. (218) Kang, K. H.; Hong, U. G.; Bang, Y. J.; Choi, J. H.; Kim, J. K.; Lee, J. K.; Han, S. J.; Song, I. K. Hydrogenation of succinic acid to 1,4butanediol over Re-Ru bimetallic catalysts supported on mesoporous carbon. Appl. Catal., A 2015, 490, 153−162. (219) Bechthold, I.; Bretz, K.; Kabasci, S.; Kopitzky, R.; Springer, A. Succinic Acid: A new platform chemical for biobased polymers from renewable resources. Chem. Eng. Technol. 2008, 31, 647−654. (220) Huang, J.; Wang, Y.; Zheng, J. M.; Dai, W. L.; Fan, K. N. Influence of support surface basicity and gold particle size on catalytic activity of Au/γ-AlOOH and Au/γ-Al2O3 catalyst in aerobic oxidation of α,ω-diols to lactones. Appl. Catal., B 2011, 103, 343−350. (221) Hou, W. B.; Dehm, N. A.; Scott, R. W. J. Alcohol oxidations in aqueous solutions using Au, Pd, and bimetallic AuPd nanoparticle catalysts. J. Catal. 2008, 253, 22−27. (222) Hwang, D. W.; Kashinathan, P.; Lee, J. M.; Lee, J. H.; Lee, U.; Hwang, J. S.; Hwang, Y. K.; Chang, J. S. Production of γ-butyrolactone from biomass-derived 1,4-butanediol over novel copper-silica nanocomposite. Green Chem. 2011, 13, 1672−1675. (223) Huang, J.; Dai, W. L.; Fan, K. Remarkable support crystal phase effect in Au/FeOx catalyzed oxidation of 1,4-butanediol to γbutyrolactone. J. Catal. 2009, 266, 228−235. (224) Yu, X. Q.; Wen, Y. Q.; Yuan, T.; Li, G. Effective production of 2,5-dimethylfuran from biomass-derived 5-hydroxymethylfurfural on ZrO2-doped graphite electrode. ChemistrySelect 2017, 2, 1237−1240. (225) Nilges, P.; Schrö der, U. Electrochemistry for biofuel generation: Production of furans by electrocatalytic hydrogenation of furfurals. Energy Environ. Sci. 2013, 6, 2925−2931. (226) Zeng, X. F.; Borole, A. P.; Pavlostathis, S. G. Biotransformation of furanic and phenolic compounds with hydrogen gas production in a microbial electrolysis cell. Environ. Sci. Technol. 2015, 49, 13667− 13675. (227) Kwon, Y.; de Jong, E.; Raoufmoghaddam, S.; Koper, M. T. Electrocatalytic hydrogenation of 5-hydroxymethylfurfural in the absence and presence of glucose. ChemSusChem 2013, 6, 1659−1667. (228) Kwon, Y.; Birdja, Y. Y.; Raoufmoghaddam, S.; Koper, M. T. Electrocatalytic hydrogenation of 5-hydroxymethylfurfural in acidic solution. ChemSusChem 2015, 8, 1745−1751. (229) Kwon, Y.; Schouten, K. J. P.; van der Waal, J. C.; de Jong, E.; Koper, M. T. M. Electrocatalytic conversion of furanic compounds. ACS Catal. 2016, 6, 6704−6717. (230) Roylance, J. J.; Kim, T. W.; Choi, K. S. Efficient and selective electrochemical and photoelectrochemical reduction of 5-hydroxymethylfurfural to 2,5-bis(hydroxymethyl)furan using water as the hydrogen source. ACS Catal. 2016, 6, 1840−1847. (231) Guo, Y. Y.; Chen, J. Z. Photo-induced reduction of biomassderived 5-hydroxymethylfurfural using graphitic carbon nitride supported metal catalysts. RSC Adv. 2016, 6, 101968−101973.

(232) Davis, S. E.; Houk, L. R.; Tamargo, E. C.; Datye, A. K.; Davis, R. J. Oxidation of 5-hydroxymethylfurfural over supported Pt, Pd and Au catalysts. Catal. Today 2011, 160, 55−60. (233) Davis, S. E.; Zope, B. N.; Davis, R. J. On the mechanism of selective oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid over supported Pt and Au catalysts. Green Chem. 2012, 14, 143− 147. (234) Gorbanev, Y. Y.; Klitgaard, S. K.; Woodley, J. M.; Christensen, C. H.; Riisager, A. Gold-catalyzed aerobic oxidation of 5hydroxymethylfurfural in water at ambient temperature. ChemSusChem 2009, 2, 672−675. (235) Martin, R. J. L. The mechanism of the Cannizzaro reaction of formaldehyde. Aust. J. Chem. 1954, 7, 335−347. (236) Swain, C. G.; Powell, A. L.; Sheppard, W. A.; Morgan, C. R. Mechanism of the Cannizzaro reaction. J. Am. Chem. Soc. 1979, 101, 3576−3583. (237) Ashby, E. C.; Coleman, D.; Gamasa, M. Single-electron transfer in the Cannizzaro reaction. J. Org. Chem. 1987, 52, 4079− 4085. (238) Blanksma, J. J. Sur le 2,5-di-oxyméthyl-furfurane. Recl. Trav. Chim. Pays-Bas Belg. 1910, 29, 403−406. (239) Middendorp, J. A. Sur l’oxyméthylfurfurol. Recl. Trav. Chim. Pays-Bas Belg. 1919, 38, 42−47. (240) Kang, E. S.; Chae, D. W.; Kim, B.; Kim, Y. G. Efficient preparation of DHMF and HMFA from biomass-derived HMF via a Cannizzaro reaction in ionic liquids. J. Ind. Eng. Chem. 2012, 18, 174− 177. (241) Subbiah, S.; Simeonov, S. P.; Esperança, J. M. S. S.; Rebelo, L. P. N.; Afonso, C. A. M. Direct transformation of 5-hydroxymethylfurfural to the building blocks 2,5-dihydroxymethylfurfural (DHMF) and 5-hydroxymethyl furanoic acid (HMFA) via Cannizzaro reaction. Green Chem. 2013, 15, 2849−2853. (242) Li, G.; Sun, Z.; Yan, Y. E.; Zhang, Y. H.; Tang, Y. Direct transformation of HMF into 2,5-diformylfuran and 2,5-dihydroxymethylfuran without an external oxidant or reductant. ChemSusChem 2017, 10, 494−498. (243) Wohlgemuth, R. Biocatalysis: key to sustainable industrial chemistry. Curr. Opin. Biotechnol. 2010, 21, 713−724. (244) Woodley, J. M. New opportunities for biocatalysis: making pharmaceutical processes greener. Trends Biotechnol. 2008, 26, 321− 327. (245) McKenna, S. M.; Leimkühler, S.; Herter, S.; Turner, N. J.; Carnell, A. J. Enzyme cascade reactions: Synthesis of furandicarboxylic acid (FDCA) and carboxylic acids using oxidases in tandem. Green Chem. 2015, 17, 3271−3275. (246) Dominguez de Maria, P.; Guajardo, N. V. Biocatalytic valorization of furans: Opportunities for inherently unstable substrates. ChemSusChem 2017, 10, 4123−4134. (247) He, Y. C.; Ding, Y.; Ma, C. L.; Di, J. H.; Jiang, C. X.; Li, A. T. One-pot conversion of biomass-derived xylose to furfuralcohol by a chemo-enzymatic sequential acid catalyzed dehydration and bioreduction. Green Chem. 2017, 19, 3844−3850. (248) He, Y. C.; Jiang, C. X.; Chong, G. G.; Di, J. H.; Wu, Y. F.; Wang, B. Q.; Xue, X. X.; Ma, C. L. Chemical-enzymatic conversion of corncob-derived xylose to furfuralcohol by the tandem catalysis with SO42−/SnO2-kaoline and E. coli CCZU-T15 cells in toluene-water media. Bioresour. Technol. 2017, 245, 841−849. (249) Zhang, Y.; Han, B.; Ezeji, T. C. Biotransformation of furfural and 5-hydroxymethyl furfural (HMF) by Clostridium acetobutylicum ATCC 824 during butanol fermentation. New Biotechnol. 2012, 29, 345−351. (250) Mussatto, S. I.; Roberto, I. C. Alternatives for detoxification of diluted-acid lignocellulosic hydrolyzates for use in fermentative processes: A review. Bioresour. Technol. 2004, 93, 1−10. (251) Palmqvist, E.; Hahn-Hagerdal, B. Fermentation of lignocellulosic hydrolysates II: Inhibitors and mechanisms of inhibition. Bioresour. Technol. 2000, 74, 25−33. (252) Liu, Z. L.; Slininger, P. J.; Dien, B. S.; Berhow, M. A.; Kurtzman, C. P.; Gorsich, S. W. Adaptive response of yeasts to furfural 2978

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis and 5-hydroxymethylfurfural and new chemical evidence for HMF conversion to 2,5-bis-hydroxymethylfuran. J. Ind. Microbiol. Biotechnol. 2004, 31, 345−352. (253) Li, Y. M.; Zhang, X. Y.; Li, N.; Xu, P.; Lou, W. Y.; Zong, M. H. Biocatalytic reduction of HMF to 2,5-bis(hydroxymethyl)furan by HMF-tolerant whole cells. ChemSusChem 2017, 10, 372−378. (254) He, Y. C.; Jiang, C. X.; Chong, G. G.; Di, J. H.; Ma, C. L. Biological synthesis of 2,5-bis(hydroxymethyl)furan from biomassderived 5-hydroxymethylfurfural by E. coli CCZU-K14 whole cells. Bioresour. Technol. 2018, 247, 1215−1220. (255) Qian, Y.; Zhu, L. F.; Wang, Y.; Lu, X. C. Recent progress in the development of biofuel 2,5-dimethylfuran. Renewable Sustainable Energy Rev. 2015, 41, 633−646. (256) Ding, J.; Zhao, J. Q.; Cheng, S. B.; Mo, X. H.; Zong, B. N. Advances in production of biobased 1,6-HDO. Chem. Ind. Eng. Prog. 2015, 34, 4209−4213. (257) Al Baradii, A.; Kokoh, K. B.; Huser, H.; Lamy, C.; Léger, J. M. Selective electrocatalytic oxidation of 2,5-dihydroxymethylfuran in aqueous medium: A chromatographic analysis of the reaction products. Electrochim. Acta 1999, 44, 2779−2787. (258) Cho, B. S.; Kim, M. J.; Jung, S. K.; Kang, S. C. Thermal decomposition kinetics and characterization of poly(butylene 2,5furandicarboxylate)/cloisite 30B composites. Korean J. Chem. Eng. 2016, 33, 3267−3272. (259) Dijkman, W. P.; Groothuis, D. E.; Fraaije, M. W. Enzymecatalyzed oxidation of 5-hydroxymethylfurfural to furan-2,5-dicarboxylic acid. Angew. Chem., Int. Ed. 2014, 53, 6515−6518. (260) Kokoh, K. B.; Belgsir, E. M. Electrosynthesis of furan-2,5dicarbaldehyde by programmed potential electrolysis. Tetrahedron Lett. 2002, 43, 229−231. (261) Seo, K. J.; Kim, M. J.; Jeong, J. H.; Lee, Y. C.; Noh, S. T.; Chung, Y. S. Polymerization and characterization of polyesters using furan monomers from biomass. Polymer (Korea) 2011, 35, 526−530. (262) Sacia, E. R.; Balakrishnan, M.; Bell, A. T. Biomass conversion to diesel via the etherification of furanyl alcohols catalyzed by Amberlyst-15. J. Catal. 2014, 313, 70−79. (263) Han, J.; Kim, J.; Jung, B.; Hwang, S.; Jegal, J.; Kim, Y. H.; Lee, Y.-S. Highly selective catalytic hydrogenation and etherification of 5hydroxymethyl-2-furaldehyde to 2,5-bis(alkoxymethyl)furans for potential biodiesel production. Synlett 2017, 28, 2299−2302. (264) Luo, J.; Yu, J. Y.; Gorte, R. J.; Mahmoud, E.; Vlachos, D. G.; Smith, M. A. The effect of oxide acidity on HMF etherification. Catal. Sci. Technol. 2014, 4, 3074−3081. (265) Shinde, S.; Rode, C. V. Cascade reductive-etherification of bioderived aldehydes over Zr-based catalysts. ChemSusChem 2017, 10, 4090−4101. (266) Jae, J.; Mahmoud, E.; Lobo, R. F.; Vlachos, D. G. Cascade of liquid-phase catalytic transfer hydrogenation and etherification of 5hydroxymethylfurfural to potential biodiesel components over Lewis acid zeolites. ChemCatChem 2014, 6, 508−513. (267) Lewis, J. D.; van de Vyver, S.; Crisci, A. J.; Gunther, W. R.; Michaelis, V. K.; Griffin, R. G.; Roman-Leshkov, Y. A continuous flow strategy for the coupled transfer hydrogenation and etherification of 5(hydroxymethyl)furfural using Lewis acid zeolites. ChemSusChem 2014, 7, 2255−2265. (268) Chen, P. X.; Tang, Y.; Zhang, B.; Liu, R. H.; Marcone, M. F.; Li, X. H.; Tsao, R. 5-Hydroxymethyl-2-furfural and derivatives formed during acid hydrolysis of conjugated and bound phenolics in plant foods and the effects on phenolic content and antioxidant capacity. J. Agric. Food Chem. 2014, 62, 4754−4761. (269) Bicker, M.; Kaiser, D.; Ott, L.; Vogel, H. Dehydration of fructose to hydroxymethylfurfural in sub- and supercritical fluids. J. Supercrit. Fluids 2005, 36, 118−126. (270) Zhu, H.; Cao, Q.; Li, C. H.; Mu, X. D. Acidic resin-catalysed conversion of fructose into furan derivatives in low boiling point solvents. Carbohydr. Res. 2011, 346, 2016−2018. (271) Liu, R. L.; Chen, J. Z.; Huang, X.; Chen, L. M.; Ma, L. L.; Li, X. J. Conversion of fructose into 5-hydroxymethylfurfural and alkyl

levulinates catalyzed by sulfonic acid-functionalized carbon materials. Green Chem. 2013, 15, 2895−2903. (272) Li, H.; Govind, K. S.; Kotni, R.; Shunmugavel, S.; Riisager, A.; Yang, S. Direct catalytic transformation of carbohydrates into 5ethoxymethylfurfural with acid-base bifunctional hybrid nanospheres. Energy Convers. Manage. 2014, 88, 1245−1251. (273) Yin, S. S.; Sun, J.; Liu, B.; Zhang, Z. H. Magnetic material grafted cross-linked imidazolium based polyionic liquids: An efficient acid catalyst for the synthesis of promising liquid fuel 5ethoxymethylfurfural from carbohydrates. J. Mater. Chem. A 2015, 3, 4992−4999. (274) Wang, Z. H.; Chen, Q. W. Conversion of 5-hydroxymethylfurfural into 5-ethoxymethylfurfural and ethyl levulinate catalyzed by MOF-based heteropolyacid materials. Green Chem. 2016, 18, 5884− 5889. (275) Xiang, B.; Wang, Y.; Qi, T.; Yang, H. Q.; Hu, C. W. Promotion catalytic role of ethanol on Brønsted acid for the sequential dehydration-etherification of fructose to 5-ethoxymethylfurfural. J. Catal. 2017, 352, 586−598. (276) Salminen, E.; Kumar, N.; Virtanen, P.; Tenho, M.; MäkiArvela, P.; Mikkola, J. P. Etherification of 5-hydroxymethylfurfural to a biodiesel component over ionic liquid modified zeolites. Top. Catal. 2013, 56, 765−769. (277) Arias, K. S.; Climent, M. J.; Corma, A.; Iborra, S. Biomassderived chemicals: Synthesis of biodegradable surfactant ether molecules from hydroxymethylfurfural. ChemSusChem 2014, 7, 210− 220. (278) Yang, F. F.; Zhang, S. G.; Zhang, Z. C.; Mao, J. B.; Li, S. M.; Yin, J. M.; Zhou, J. X. A biodiesel additive: etherification of 5hydroxymethylfurfural with isobutene to tert-butoxymethylfurfural. Catal. Sci. Technol. 2015, 5, 4602−4612. (279) Sousa, A. F.; Vilela, C.; Fonseca, A. C.; Matos, M.; Freire, C. S. R.; Gruter, G. J. M.; Coelho, J. F. J.; Silvestre, A. J. D. Biobased polyesters and other polymers from 2,5-furandicarboxylic acid: A tribute to furan excellency. Polym. Chem. 2015, 6, 5961−5983. (280) Papageorgiou, G. Z.; Papageorgiou, D. G.; Terzopoulou, Z.; Bikiaris, D. N. Production of bio-based 2,5-furan dicarboxylate polyesters: Recent progress and critical aspects in their synthesis and thermal properties. Eur. Polym. J. 2016, 83, 202−229. (281) Zhang, J. H.; Li, J. K.; Tang, Y. J.; Lin, L.; Long, M. N. Advances in catalytic production of bio-based polyester monomer 2,5furandicarboxylic acid derived from lignocellulosic biomass. Carbohydr. Polym. 2015, 130, 420−428. (282) Xu, S.; Zhou, P.; Zhang, Z. H.; Yang, C. J.; Zhang, B. G.; Deng, K. J.; Bottle, S.; Zhu, H. Y. Selective oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid using O2 and a photocatalyst of co-thioporphyrazine bonded to g-C3N4. J. Am. Chem. Soc. 2017, 139, 14775−14782. (283) Lei, D.; Yu, K.; Li, M. R.; Wang, Y. L.; Wang, Q.; Liu, T.; Liu, P. K.; Lou, L. L.; Wang, G. C.; Liu, S. X. Facet effect of singlecrystalline Pd nanocrystals for aerobic oxidation of 5-hydroxymethyl-2furfural. ACS Catal. 2017, 7, 421−432. (284) Moore, J. A.; Kelly, J. E. Polyesters derived from furan and tetrahydrofuran nuclei. Macromolecules 1978, 11, 568−573. (285) Fawcett, A. H.; Yau, T. F.; Mulemwa, J. N.; Tan, C. E. The free-radically prepared copolymers of acrylonitrile with furfuryl alcohol and similar furan derivatives. Br. Polym. J. 1987, 19, 211−221. (286) Boufi, S.; Belgacem, M. N.; Quillerou, J.; Gandini, A. Urethanes and polyurethanes bearing furan moieties 4. Synthesis, kinetics and characterization of linear polymers. Macromolecules 1993, 26, 6706−6717. (287) Boufi, S.; Gandini, A.; Belgacem, M. N. Urethanes and polyurethanes bearing furan moieties: 5. Thermoplastic elastomers based on sequenced structures. Polymer 1995, 36, 1689−1696. (288) Boufi, S.; Belgacem, M. N.; Gandini, A. 2-Furyloxiranes. III. Chain extension with different polyols. Polym. J. (Tokyo, Jpn.) 1997, 29, 479−486. 2979

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980

Review

ACS Catalysis (289) Laita, H.; Boufi, S.; Gandini, A. The application of the Diels− Alder reaction to polymers bearing furan moieties. 1. Reactions with maleimides. Eur. Polym. J. 1997, 33, 1203−1211. (290) Zhang, Y.; Li, T.; Xie, Z. N.; Han, J. R.; Xu, J.; Guo, B. H. Synthesis and properties of biobased multiblock polyesters containing poly(2,5-furandimethylene succinate) and poly(butylene succinate) blocks. Ind. Eng. Chem. Res. 2017, 56, 3937−3946. (291) Lillie, L. M.; Tolman, W. B.; Reineke, T. M. Structure/ property relationships in copolymers comprising renewable isosorbide, glucarodilactone, and 2,5-bis(hydroxymethyl)furan subunits. Polym. Chem. 2017, 8, 3746−3754. (292) Zeng, C.; Seino, H.; Ren, J.; Hatanaka, K.; Yoshie, N. Biobased furan polymers with self-healing ability. Macromolecules 2013, 46, 1794−1802. (293) Zeng, C.; Seino, H.; Ren, J.; Hatanaka, K.; Yoshie, N. Selfhealing bio-based furan polymers cross-linked with various bismaleimides. Polymer 2013, 54, 5351−5357. (294) Ikezaki, T.; Matsuoka, R.; Hatanaka, K.; Yoshie, N. Biobased poly(2,5-furandimethylene succinate-co-butylene succinate) crosslinked by reversible Diels−Alder reaction. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 216−222. (295) Zeng, C.; Seino, H.; Ren, J.; Yoshie, N. Polymers with multishape memory controlled by local glass transition temperature. ACS Appl. Mater. Interfaces 2014, 6, 2753−2758. (296) Jiang, Y.; Woortman, A. J.; Alberda van Ekenstein, G. O.; Petrovic, D. M.; Loos, K. Enzymatic synthesis of biobased polyesters using 2,5-bis(hydroxymethyl)furan as the building block. Biomacromolecules 2014, 15, 2482−2493. (297) Azim, H.; Dekhterman, A.; Jiang, Z.; Gross, R. A. Candida antarctica Lipase B-catalyzed synthesis of poly(butylene succinate): Shorter chain building blocks also work. Biomacromolecules 2006, 7, 3093−3097. (298) Juais, D.; Naves, A. F.; Li, C.; Gross, R. A.; Catalani, L. H. Isosorbide polyesters from enzymatic catalysis. Macromolecules 2010, 43, 10315−10319. (299) Jiang, Z. Lipase-catalyzed copolymerization of dialkyl carbonate with 1,4-butanediol and omega-pentadecalactone: Synthesis of poly(omega-pentadecalactone-co-butylene-co-carbonate). Biomacromolecules 2011, 12, 1912−1919. (300) Kobayashi, T.; Matsumura, S. Enzymatic synthesis and properties of novel biodegradable and biobased thermoplastic elastomers. Polym. Degrad. Stab. 2011, 96, 2071−2079. (301) Jiang, Y.; Woortman, A. J.; van Ekenstein, G. O.; Loos, K. Enzyme-catalyzed synthesis of unsaturated aliphatic polyesters based on green monomers from renewable resources. Biomolecules 2013, 3, 461−480. (302) Liu, W.; Wang, F.; Tan, T.; Chen, B. Lipase-catalyzed synthesis and characterization of polymers by cyclodextrin as support architecture. Carbohydr. Polym. 2013, 92, 633−640. (303) Zhang, Y. R.; Spinella, S.; Xie, W. C.; Cai, J. L.; Yang, Y. X.; Wang, Y. Z.; Gross, R. A. Polymeric triglyceride analogs prepared by enzyme-catalyzed condensation polymerization. Eur. Polym. J. 2013, 49, 793−803. (304) Harris, J. M.; Keranen, M. D.; O’Doherty, G. A. Syntheses of Dand L-mannose, gulose, and talose via diastereoselective and enantioselective dihydroxylation reactions. J. Org. Chem. 1999, 64, 2982−2983. (305) Martin, S. F.; Chen, H. J.; Yang, C. P. Facile asymmetric syntheses of L-deoxycastanospermine and L-deoxy-8a-epi-castanospermine. J. Org. Chem. 1993, 58, 2867−2873. (306) Wender, P. A.; Rice, K. D.; Schnute, M. E. The first formal asymmetric synthesis of phorbol. J. Am. Chem. Soc. 1997, 119, 7897− 7898. (307) Noutsias, D.; Alexopoulou, I.; Montagnon, T.; Vassilikogiannakis, G. Using water, light, air and spirulina to access a wide variety of polyoxygenated compounds. Green Chem. 2012, 14, 601−604. (308) Horvat, J.; Klaić, B.; Metelko, B.; Šunjić, V. Mechanism of levulinic acid formation. Tetrahedron Lett. 1985, 26, 2111−2114.

(309) Camblor, M. A.; Corma, A.; Martínez, A.; Pérez-Pariente, J. Synthesis of a titaniumsilicoaluminate isomorphous to zeolite beta and its application as a catalyst for the selective oxidation of large organic molecules. J. Chem. Soc., Chem. Commun. 1992, 0, 589−590. (310) Corma, A.; Navarro, M. T.; Pariente, J. P. Synthesis of an ultralarge pore titanium silicate isomorphous to MCM-41 and its application as a catalyst for selective oxidation of hydrocarbons. J. Chem. Soc., Chem. Commun. 1994, 147−148. (311) Wahlen, J.; Moens, B.; de Vos, D. E.; Alsters, P. L.; Jacobs, P. A. Titanium silicalite 1 (TS-1) catalyzed oxidative transformations of furan derivatives with hydrogen peroxide. Adv. Synth. Catal. 2004, 346, 333−338. (312) Noutsias, D.; Kouridaki, A.; Vassilikogiannakis, G. Scope and limitations of the photooxidations of 2-(α-hydroxyalkyl)furans: Synthesis of 2-hydroxy-exo-brevicomin. Org. Lett. 2011, 13, 1166− 1169.

2980

DOI: 10.1021/acscatal.7b03530 ACS Catal. 2018, 8, 2959−2980