Catalytic Advances in the Production and Application of Biomass

Feb 28, 2018 - Namely, the adsorbed hydrogen (Hads) on the metal surface was unlikely to be involved in the formation of the target product, and there...
12 downloads 9 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

Review

Catalytic advances in the production and application of biomass-derived 2,5-dihydroxymethylfuran Lei Hu, Jiaxing Xu, Shouyong Zhou, Aiyong He, Xing Tang, Lu Lin, Jiming Xu, and Yijiang Zhao ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b03530 • Publication Date (Web): 28 Feb 2018 Downloaded from http://pubs.acs.org on February 28, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Catalysis is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Catalytic advances in the production and application of biomass-derived 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

*Corresponding Author: [email protected] Telephone/Fax: +86-0517-83526983

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 due to its peculiar symmetrical structure and wide potential applications. At present, studies on the production of DHMF are quickly progressing, with productive approaches being increasingly developed, many crucial achievements have been continually obtained. However, to date, a special and real-time review of this research area is still lacking. To

1

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 95

gain more insight into the current research situation, this review comprehensively summarizes and discusses state-of-the-art advancements of the production of DHMF from HMF via various chemocatalytic pathways, such as conventional hydrogenation, transfer hydrogenation, electrocatalytic hydrogenation, photocatalytic hydrogenation and disproportionation reaction, and biocatalytic pathways. Meanwhile, this review also systematically outlines the latest results of 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

1. INTRODUCTION At present, most chemicals and fuels in the world are directly or indirectly produced from the non-renewable fossil resources.1-5 However, as the depletion of fossil resources, exploring appropriate renewable resources has becoming increasingly urgent and necessary.6-10 As the only carbon-based 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, biorefinery is a very momentous approach.17-20 During the biorefinery 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 starch49-54 and cellulose,55-60 is hailed as one of the most

2

ACS Paragon Plus Environment

Page 3 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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),66-70

2,5-dihydroxymethyltetrahydrofuran (DHMTHF),71-76 2,5-dimethylfuran (DMF),77-88 2,5-dimethyltetrahydrofuran (DMTHF),89-92 1-hydroxyhexane-2,5-dione (HHD),93-99 1,2,6-hexanetriol

(HTO),100-102

1,6-hexanediol

(HDO)103-107

and

3-hydroxymethylcyclopetanone (HMCPN).108-110 Among them, DHMF is highly attractive due to its peculiar symmetrical structure and wide potential applications in the production of ethers,111-113 ketones114 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 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 and 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 the available catalytic pathways for the transformation of biomass-derived fundamental compounds,

3

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 95

such as glycerol (GC), furfural (FF) and levulinic acid (LA), in which the selective hydrogenation of HMF into DHMF was 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 were 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 were also systematically outlined. In a word, the main aim of this review is to attract concerns about DHMF and provide some theoretical references and technical supports for the practical production and application of DHMF in the near future.

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 reductive agents and the co-production 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 section, we summarized various chemocatalytic pathways

including

conventional

hydrogenation,

transfer

hydrogenation,

electrocatalytic hydrogenation, photocatalytic hydrogenation and disproportionation

4

ACS Paragon Plus Environment

Page 5 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

reaction in a comprehensive manner, and we also discussed 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 hydrogen donor for the selective hydrogenation of HMF is the most common and important pathway to produce DHMF, which has been widely studied over various catalysts in recent years. For example, ruthenium-based dinuclear complex (Ru-based Shvo’s catalyst) was used by Pasini et al. in 2014 for the selective hydrogenation of HMF.138 Under 10 bar H2, an excellent yield of DHMF of 99% 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 played an important role in the catalytic cycle towards 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 a Ru-based Shvo’s 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).

5

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 95

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 with 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, 82∼100% yields of DHMF could be obtained 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 heterogeneous catalysts are composed of active metals and auxiliary supports. Apart from the metal properties, such as electron deficiency, particle size and dispersion

6

ACS Paragon Plus Environment

Page 7 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

degree, the selective hydrogenation of HMF into DHMF may be strongly affected by the natures of supports. To verify this speculation, Ohyama et al. prepared a variety of gold catalysts.148 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 DHMF was formed under the same reaction conditions, suggesting that compared with acidic supports, basic supports were more conducive to the formation of DHMF, which was probably caused by favoring 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 between active metals and auxiliary supports. Among the above heterogeneous catalysts, whether Ru, Pt, Ir or Au, they are precious metals, and therefore, are very expensive. From the viewpoint of economization and industrialization, the development of non-precious metal-based heterogeneous catalysts are highly desirable for the selective hydrogenation of HMF into DHMF. To explore the potential of non-precious 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,

7

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 95

ethanol, 1,4-dioxane and water, the yields of DHMF were up to 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 non-precious metal-based heterogeneous catalysts, especially the Cu-based heterogeneous catalysts, possessed similar catalytic performance to that of precious metal-based heterogeneous catalysts for the selective hydrogenation of HMF into DHMF. More gratifyingly, by adjusting the hydrogen pressure and reaction temperature, these non-precious metal-based heterogeneous catalysts could also 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) that was developed by Yu et al. provided a 96.1% yield of DHMF at 110 °C for 18 h in the presence of 30 bar H2, whereas DHMF yield was only 76.4% over carbon nanotube-supported

nickel

(Ni/CNT).153

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 calculational results of density functional theory (DFT). 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 the η2-(C,O)-coordination, which

8

ACS Paragon Plus Environment

Page 9 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

would further promote the selective hydrogenation of C=O of HMF on the bimetallic alloy due to a 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; the 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 more milder reaction conditions for the

selective

hydrogenation of

HMF

into

DHMF.157-159 Particularly, 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 could be 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 types of catalysts, the selective hydrogenation of HMF into DHMF is also highly associated with the properties of solvents. Generally, the employed solvents in the selective hydrogenation of HMF into DHMF can be divided into polar protic solvents (such as water, methanol, ethanol, propanol and butanol), polar aprotic solvents (such as acetone and tetrahydrofuran) and nonpolar solvents (such as hexane), and they have various δ values that represent the difference between the acceptor and donor numbers, which are related to a quantitative measure of the Lewis acidity and basicity, respectively. Due to the negative δ values, the protic solvents that are capable of accepting electrons showed better performance than the aprotic

9

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

solvents with positive δ values that are the symbol of donating electrons, which 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 values of solvents had a great impact on the conversion of substrates.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 a harder polarization of C=O in nonpolar solvents.140 Additionally, the pH values of solutions are also a momentous factor influencing the selective hydrogenation of HMF into DHMF. In contrast to a solution of pH = 7, when this reaction was conducted in a solution of pH < 7, the selectivity of DHMF was unsatisfactory even though HMF could be nearly converted 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 acid-catalyzed ring-opening and subsequent hydrogenation of DHMF.141 If this reaction was conducted in a solution of 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 neutral water with a smaller δ value and a stronger polarity is regarded as a good solvent, which has 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

10

ACS Paragon Plus Environment

Page 10 of 95

Page 11 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

DHMF, in which H2 is mostly employed as a hydrogen donor owing to 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 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 form of liquid at 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, for alcohols, they can also act as reaction mediums. Hence, when alcohols are used as hydrogen donors, additional reaction mediums are no longer needed, which can also enhance the economy of CTH to some extent. If alcohols are cheap enough, the economy of CTH

11

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 95

will be more obvious. More importantly, after the end of CTH, the unconverted alcohols can be separated from the reaction mixture and then reused as hydrogen donors and reaction mediums, and the converted alcohols, that is, the produced aldehydes or ketones, can also be separated and then integrated into the other reaction steps, such as carbon chain growth reaction via aldol condensation. Based on 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 LA184-188 and ethyl levulinate (EL),189-193 via the reaction of MPV. 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 co-product in the acid-catalyzed decomposition of biomass-derived carbohydrates into LA,194-198 was firstly adopted as a hydrogen donor by Thananatthanachon and Rauchfuss.199 Surprisingly, the yield of DHMF 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,

12

ACS Paragon Plus Environment

Page 13 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Pd/C was found to be capable of tolerating formic acid, and when it was used as a heterogeneous catalyst for the selective hydrogenation of HMF, a comparable yield of DHMF 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 H2-based hydrogenation process, but it has still been found to be an effective pathway, which can proceed through the 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, a 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 were proved to play a synergistic role in the direct hydrogen transfer (Figure 6). Enlightened by the acid-base bifunctional actions of ZrO(OH)2, a series of zirconium-based organic-inorganic coordination

13

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

polymers

such

as

zirconium-alkyltriphosphate

Page 14 of 95

nanohybrid

(Zr-ATPN),

zirconium-benzylphosphonate nanohybrid (Zr-BPPN) and zirconium-furandicarboxylate nanohybrid (Zr-FDCN) were designed by Li et al. for the selective hydrogenation of HMF.204-206 In the presence of isopropanol, DHMF yields 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 hydroxyapatite-encapsulated ferric oxide (Fe2O3@HAP),207 mesoporous carbon-supported cobaltosic oxide (Co3O4@MC),208 silicon carbide-supported gold (Au/SiC),209 cobaltosic oxide-supported ruthenium (Ru/Co3O4)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 possessed a broad universality to 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 (∆Hf°) is defined as the difference of the standard molar enthalpy of formation between an alcohol and its corresponding carbonyl compound, and it represents the complexity of hydrogen abstraction. As shown in Table 3, the reduction potentials of various alcohols decrease in the order of methanol > ethanol > 1-butanol > isopropanol ≈ 2-butanol. Compared with primary alcohols, secondary alcohols with lower reduction potentials are more appropriate as hydrogen donors in the process of CTH, and this may be the reason for

14

ACS Paragon Plus Environment

Page 15 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

that isopropanol is widely used for the selective hydrogenation of HMF into DHMF. Due to similar reduction potential with 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 so, it does not easily act as a hydrogen donor in theory. Unexpectedly, Pasini et al. 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).216 After 3 h at 160 °C, almost 100% yield of DHMF was achieved, and more surprisingly, the co-products were exclusively gaseous compounds, such as carbon monoxide (CO), carbon dioxide (CO2) and methane (CH4). From experimental results and theoretical calculations, the further study of reaction mechanism revealed that the resulting formaldehyde could react with the remaining methanol to produce hemiacetal, which could also act as a hydrogen donor for the selective hydrogenation of HMF into DHMF (Figure 7). Meanwhile, its dehydrogenated product, methylformate, would further decompose into the abovementioned gaseous co-products, which were easily removed in the reactor depressurization process.216


15

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In addition to the common alcohols, benzyl alcohol was also tested by Gao et al. 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.171 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% could be achieved in only 0.6 minute at 220 °C.137 Notably, one molecule of BDO can provide two molecules of H2 and can then be converted into γ-butyrolactone (GBL) by the oxygen-free dehydrogenation and ring-closure reaction, which is also a versatile intermediate for the synthesis of fine chemicals.220-223 2.3. Electrocatalytic hydrogenation pathway The 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. firstly studied the selective hydrogenation of HMF via an electrocatalytic pathway in 0.1 M sodium sulfate (Na2SO4).227 By correlating 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 performance, leading to an 85% selectivity at

16

ACS Paragon Plus Environment

Page 16 of 95

Page 17 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

−0.81 V. However, if the selective hydrogenation of HMF was conducted in 0.5 M sulfuric acid (H2SO4),228 the most effective electrode was nickel (Ni), indicating that the electrocatalytic hydrogenation of HMF into DHMF was strongly influenced by the types of electrodes and electrolytes, which could be 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 electrolytes,227-229 which is shown in Figure 9. In a neutral solution (pH = 7),227 the onset potentials of the target product formation on all metal electrodes were −0.5 ± 0.2 V, which were less negative than those of the HER, ranging 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 potentials, ranging from −0.4 to 0 V on transiƟon 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 onset potentials of HER. In both neutral and acidic solutions, the HER could occur via the generation of Hads,

17

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

regardless of whether its source was H2O or H+: H+ + e− ↔ Hads or H2O + e− ↔ Hads + OH−. Interestingly, HER in the acidic solution was much more effective than in the neutral solution, suggesting that Hads on the metal surface was involved in the formation of the target product in an 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 et al.230 As shown in Figure 10, the formation of DHMF could proceed through the 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 et al. also constructed an ideal photoelectrocatalytic pathway by simply replacing a metal anode (such as Pt) in electrocatalytic pathway with a photoanode (such as BiVO4).230 Under illumination, BiVO4 could generate 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 a 95% DHMF selectivity and a 94% faradaic efficiency (FE).230 This photoelectrocatalytic pathway provides a practical, inexpensive and 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.

18

ACS Paragon Plus Environment

Page 18 of 95

Page 19 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Note that the electrocatalytic pathway has been deeply studied, but 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 the single solar cell. 2.4. Photocatalytic hydrogenation pathway Inspired by the photoelectrocatalytic pathway, a pure photocatalytic pathway was developed by Guo and Chen in 2016 for the selective hydrogenation of HMF into DHMF,231 in which graphitic carbon nitride-supported platinum (Pt/g-C3N4) was

19

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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 Pt/g-C3N4 could not only facilitate the photoinduced water splitting to produce hydrogen but also promote the subsequent activation of the produced hydrogen for the selective hydrogenation of HMF, which should be due to the synergistic action of Pt and g-C3N4.231 Under visible light irradiation of a 210 W xenon lamp equipped with a 420 nm cut-off 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 time for the direct photocatalytic hydrogenation of HMF into DHMF, and most importantly, 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, is far from large-scale applications. On this occasion, how to design and prepare a more effective photocatalyst and balance the roles of 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 abovementioned 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

20

ACS Paragon Plus Environment

Page 20 of 95

Page 21 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

decades, only two reports on the Cannizzaro reaction of HMF were published by Blanksma238 and Middendorp239 in 1910 and 1919, respectively. Excitingly, in 2012 and 2013, Kang et al.240 and Subbiah et al.241 restudied the Cannizzaro reaction of HMF, respectively. More importantly, the outstanding yields of DHMF 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 with 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. More recently, inspired by the Cannizzaro reaction of HMF, a novel acid-catalyzed disproportionation reaction over trimethylaluminum (AlMe3) was developed in acetonitrile by Li et al. for the selective hydrogenation of HMF into DHMF,242 in which 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

21

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

formed in the presence of AlMe3, HMF conversion was only 26.7% at 80 °C for 1 h. By combining 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 a poor solubility in acetonitrile, and therefore, its precipitation from the reaction mixture should shift the redox equilibrium into the desired products. Based on this speculation, a cooling treatment in an ice-water bath at an interval of ten minutes was applied in the acid-catalyzed disproportionation reaction of HMF. Surprisingly, with this strategy, 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 to the production and purification of DHMF.

3. BIOCATALYTIC HYDROGENATION OF HMF INTO DHMF As 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 with a number of advantages, such as excellent selectivity, high efficiency, environmental friendliness and mild reaction conditions, may be a more

22

ACS Paragon Plus Environment

Page 22 of 95

Page 23 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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. to examine their responses to HMF.252 Surprisingly, among three strains, S. cerevisiae NRRL Y-12632 could not only tolerate 30 mM HMF but also 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. in 2017, and it was proved to be more effective for the biocatalytic hydrogenation of HMF into DHMF.253 When the concentration of HMF was 100 mM, a yield of DHMF of 86% could be achieved at 35 °C for 12 h. Following this exciting result, He et al. found that recombinational Escherichia coli CCZU-K14 also displayed an amazing tolerance and catalytic ability towards HMF (Figure 15).254 When the concentration of HMF was up to 400 mM, a 70.2% yield of DHMF was still obtained at 30 °C for 72 h. If 100 and 200 mM HMF were used as substrates, DHMF yield could even be as high as 100 and 90.2% at the same reaction

23

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

temperature, respectively. Obviously, the biocatalytic hydrogenation of HMF into DHMF was successfully realized in the presence of E. coli CCZU-K14. Moreover, reduced nicotinamide adenine dinucleotide (NADH) or reduced nicotinamide adenine dinucleotide phosphate (NADPH) as a co-factor was crucial to the biocatalytic hydrogenation reaction, and its effective regeneration was closely related to the co-substrate 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 in industrial conditions. Considering the industrial production, the immobilized whole cells with excellent reusability and high productivity are highly recommended.

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, therefore, by adopting 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, in the earlier studies, 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

24

ACS Paragon Plus Environment

Page 24 of 95

Page 25 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

will be introduced in detail. 4.1. Through the etherification reaction Due to 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 5-alkoxymethylfurfural (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. firstly reported the etherification of DHMF with ethanol in 2012.112 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

2,5-bis(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 satisfied, and the yield of 2,5-bis(methoxymethyl)furan (BMMF) was only 57% at 60 °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, acidic zeolites were selected by Cao et al. to obtain a satisfactory yield of BMMF.113 Among various acidic zeolites, HZSM-5 with a Si/Al ratio of 25 was the most effective

25

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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. Note that the smaller ratio of Si/Al of zeolites meant more acidic sites and stronger hydrophilicities, which facilitated the etherification reaction between DHMF and methanol on the surface of catalysts.113 Based on these results, it can be seen that the properties of catalysts and the types of alcohols have a great effect on the etherification of DHMF into BAMF. In addition, it should be specially noted that various BAMF 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 the hydrogenation catalysts and the subsequent etherification of DHMF into BAMF over the acid catalysts (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 of terephthalic acid for the synthesis of various polymers.279-283 Correspondingly, it is reasonable to assume that DHMF is also a suitable replacement of 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

26

ACS Paragon Plus Environment

Page 26 of 95

Page 27 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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 so, 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 then led to the occurrence of undesired reactions.284 Therefore, solution polymerization should be more feasible. Recently, Yoshie and co-workers reported a series of studies,292-295 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 crosslinked 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 controlling the number and concentration of M2. With this strategy, PFS/M2 displayed an excellent multi-shape memory and recovery behavior (Figure 18). Subsequently, an ultraviolet-induced photopolymerization method was developed by Jang et al. 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-8(E)-octadecenoic acid (ADOD),118 in which 2,5-furan diacrylate that 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

27

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

after only 15 min, the gel content in all network polymers was higher than 97%.118 More recently, Jiang et al. 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).296 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 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, α-6-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. via the ring-rearrangement of DHMF in the presence of water over a heterogeneous acid catalyst of Amberlyst-15.114 Notably, the same treatment in alcohol led to the etherification of hydroxy groups, and no rearrangement reaction was observed. Under the atmospheric pressure of nitrogen (N2), the yield and selectivity of HMEPO could be

28

ACS Paragon Plus Environment

Page 28 of 95

Page 29 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

achieved to 50 and 70% in 30 min at a moderate reaction temperature of 70 °C, respectively. Furthermore, this ring-rearrangement reaction does not require the addition of oxidants, which is different from the Achmatowicz rearrangement reaction that is a typical oxidative process for the conversion of 2-(α-hydroxyalkyl)furan into 6-hydroxy-3(2H)-pyranones. Based on the previous reports308-312 and the analysis results of gas chromatography mass spectrometry (GC-MS) and proton nuclear magnetic resonance spectroscopy (1H NMR), a plausible reaction mechanism for the ring-rearrangement of DHMF into HMEPO was proposed in Figure 20, which involved hydration, dehydration, ring opening, keto-enol tautomerism and ring closure. 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-2,5-dione, 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 pathways and biocatalytic pathways. From the current point of view, among these pathways, the conventional hydrogenation pathway with external

29

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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 the possible bottlenecks, the main one should be attributed to 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 concerns 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) The 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 non-precious 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 catalysts to the complex systems should also be considered. (II) The intensive exploration of appropriate reaction mediums. During the production and application of DHMF, reaction mediums can act not only as solvents but also as hydrogen donors and reactants. From the viewpoint of green chemistry, the development of inexpensive and renewable reaction mediums with strong solvencies and particular functionalities is imminently necessary. (III) The comprehensive optimization of reaction parameters. Once the exclusive catalysts and reaction mediums are chosen in the production and application of DHMF, various reaction

30

ACS Paragon Plus Environment

Page 30 of 95

Page 31 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

parameters such as reaction temperature, reaction time, catalyst amount and substrate concentration will have a more direct influence on the formation of target products. Thus, to get higher yields, they should be thoroughly optimized. (IV) The strategic establishment of separation methods. It is generally known that DHMF is not very stable especially in high-temperature circumstance and the reaction mixture is quite complicated in its production and application. In this situation, the effective separation of target products from the complicated reaction mixture is the key for their practical uses. Therefore, in addition to getting the higher yields of target products, establishing high-efficiency and energy-efficient separation methods based on the respective physicochemical properties of various target products is urgently needed. (V) The 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 the recycle of catalysts and reaction mediums and separating equipment for the separation of target products should be creatively designed according to the corresponding catalytic systems. Apart from these scientific issues, the in-depth techno-economic analysis 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 will be exceptionally bright.

ACKNOWLEDGMENTS 31

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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. Renew. Sustain. 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.

32

ACS Paragon Plus Environment

Page 32 of 95

Page 33 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(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. (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. Renew. Sustain. 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: Environ. 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.

33

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 95

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. 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.; 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.

34

ACS Paragon Plus Environment

Page 35 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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.;

Santamaría-González,

J.;

Maireles-Torres, P.; Jiménez-López, A. Selective dehydration of glucose to

35

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

5-hydroxymethylfurfural on acidic mesoporous tantalum phosphate. Appl. Catal. B: Environ. 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 5-hydroxymethylfurfural in superheated water. Bioresour. Technol. 2013, 132, 342-350. (36) Arenas-Ligioiz, I.; Pérez-Maqueda, J.; López, Ó.; Fernández-Bolaños, 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.

36

ACS Paragon Plus Environment

Page 36 of 95

Page 37 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(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. 5-hydroxymethylfurfural

Conversion of

(HMF)

by

glucose

rare

and cellobiose

earth

metal

salts

into 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-2-furaldehyde 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 5-hydroxymethylfurfural 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. Zeolite-promoted 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.

37

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Acid-catalyzed dehydration of fructose and inulin with glycerol or glycerol carbonate as renewably sourced co-solvent. 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. One-pot 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 aluminium chloride catalyst in water. Green Chem. 2011, 13, 2859-2868. (50) Nikolla, E.; Román-Leshkov, Y.; Moliner, M.; Davis, M. E. “One-pot” synthesis of 5-(hydroxymethyl)furfural from carbohydrates using tin-beta zeolite. ACS Catal. 2011, 1, 408-410. (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 Reyes, A. A.; Luque, R. Catalytic conversion of starch into valuable furan derivatives using supported metal

38

ACS Paragon Plus Environment

Page 38 of 95

Page 39 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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.; Tsang, D. C.; Yip, A. C.; 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: Environ. 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

39

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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.

40

ACS Paragon Plus Environment

Page 40 of 95

Page 41 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(67) Verevkin, S. P.; Emel'yanenko, V. N.; Stepurko, E. N.; Ralys, R. V.; Zaitsau, D. H. Biomass-derived platform chemicals: Thermodynamic studies on the conversion of 5-hydroxymethylfurfural into bulk intermediates. Ind. Eng. Chem. Res. 2009, 48, 10087-10093. (68) Scholz, D.; Aellig, C.; Hermans, I. Catalytic transfer hydrogenation/hydrogenolysis for reductive upgrading of furfural and 5-(hydroxymethyl)furfural. ChemSusChem 2013, 7, 268-275. (69)

Chatterjee,

M.;

Ishizaka,

T.;

Kawanami,

H.

Hydrogenation

of

5-hydroxymethylfurfural 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 metal-organic 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

41

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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-aza-bicyclo[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 2,5-dimethylfuran over Ni-Co/C catalyst. Fuel 2017, 187, 159-166. (78) Luo, J.; Monai, M.; Wang, C.; Lee, J. D.; Duchoň, T.; Dvořák, 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 5-hydroxymethylfurfural to 2,5-dimethylfuran over a

42

ACS Paragon Plus Environment

Page 42 of 95

Page 43 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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. Nickel-tungsten 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 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

43

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 95

(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. 2015, 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.

44

ACS Paragon Plus Environment

Page 45 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(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: Gen. 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 5-HMF 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.

45

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 95

(102) Nolan, M. R.; Sun, G.; Shanks, B. H. On the selective acid-catalysed 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-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 1,2,6-hexanetriol 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 1,6-hexanediol from HMF over double-layered catalysts of Pd/SiO2+Ir-ReOx/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 ring-rearrangement of biomass

46

ACS Paragon Plus Environment

Page 47 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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

5-hydroxymethylfurfural 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-bis-methoxymethylfuran: A promising cetane number improver for diesel. Appl. Catal. A: Gen. 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 bio-based 2,5-bis-hydroxymethylfuran into α-6-hydroxy-6-methyl-4-enyl-2H-pyran-3-one in water. Appl. Catal. B: Environ. 2016, 180, 38-43.

47

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 48 of 95

(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: Gen. 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. 2014, 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

2,5-dimethylfuran

carbon-supported ruthenium. Ind. Eng. Chem. Res. 2014, 53, 3056-3064.

48

ACS Paragon Plus Environment

over

Page 49 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(123) Hu, L.; Lin, L.; Liu, S. J. Chemoselective hydrogenation of biomass-derived 5-hydroxymethylfurfural into the liquid biofuel 2,5-dimethylfuran. 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.;

Vaidy,

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: Gen. 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: Gen. 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.

49

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(131)

Milone,

C.

Selective

hydrogenation

Page 50 of 95

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.; Brusta, 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.;

50

ACS Paragon Plus Environment

Page 51 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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. (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.

51

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 52 of 95

(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 2,5-dimethylfuran 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

52

ACS Paragon Plus Environment

Page 53 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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: Environ. 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 5-hydroxymethylfurfural. ChemCatChem 2015, 7, 1701-1707. (154) Srivastava, S.; Jadeja, G. C.; Parikh, J. Synergism studies on alumina-supported copper-nickel

catalysts

towards

furfural

and

5-hydroxymethylfurfural

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 Sustain. Chem. Eng. 2013, 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.

53

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(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.

54

ACS Paragon Plus Environment

Page 54 of 95

Page 55 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

RSC Adv. 2013, 3, 2112-2153. (166) Aramendı ́a, M. A.; Borau, V.; Jiménez, C.; Marinas, J. M. Liquid-phase heterogeneous catalytic transfer hydrogenation of citral on basic catalysts. J. Mol. Catal. A: Chem. 2001, 171, 153-158. (167) Aramendı ́a, 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: Gen. 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) López, R. M.; Maireles-Torres, P.; Ojeda, M.; Sadaba, I.; Granados, M. L. 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.

55

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 56 of 95

(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: Gen. 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: Environ. 2012, 128, 31-38. (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: Environ. 2013, 140-141, 386-395.

56

ACS Paragon Plus Environment

Page 57 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(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 zirconium-based 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.; 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

57

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

conversion of biomass-derived levulinic acid into γ-valerolactone 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 gamma-valerolactone over metal oxide catalysts. Chem. Commun. 2011, 44, 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. L.; Zhu, J. H.; Xu, X.; Wang, W. M.; Li, J. W.; Zhao, Y.; Tang, K. J.; Song, Q.; Qi, X. L.; Kong, D. J.; Tang, Y. Continuous hydrogenation of ethyl levulinate to gamma-valerolactone and 2-methyl tetrahydrofuran over alumina doped Cu/SiO2 catalyst: the potential of commercialization. Sci. Rep. 2016, 6, 28898-28906. (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: Gen. 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 γ-valerolactone over mesoporous Zr/B mixed oxides. J. Ind. Eng. Chem. 2016, 43, 133-141.

58

ACS Paragon Plus Environment

Page 58 of 95

Page 59 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(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. Renew. Sustain. 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.

59

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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 2,5-dimethylfuran. ChemCatChem 2014, 6, 848-856. (202) Nagpure, A. S.; Venugopal, A. K.; Lucas, N.; Manikandan, M.; Thirumalaiswamy, R.; Chilukuri, S. Renewable fuels from biomass-derived 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 5-hydroxymethyl 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 Zr-benzylphosphonate 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, 18, 1761-1770. (207) Wang, F.; Zhang, Z. H. Catalytic transfer hydrogenation of furfural into furfuryl alcohol over magnetic γ-Fe2O3@HAP catalyst. ACS Sustain. Chem. Eng. 2017, 5,

60

ACS Paragon Plus Environment

Page 60 of 95

Page 61 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

942-947. (208) Wang, G. H.; Deng, X. H.; Gu, D.; Chen, K.; Tîysîz, H.; Spliethoff, B.; Bongard, H. J.; Weidenthaler, C.; Schmidt, W.; Schîth, 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, 10, 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-Ponndorf-Verley-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

61

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 62 of 95

diols: Key intermediates for sustainable chemicals, materials and fuels. Renew. Sustain. 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.; 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 1,4-butanediol over Re-Ru bimetallic catalysts supported on mesoporous carbon. Appl. Catal. A: Gen. 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

62

ACS Paragon Plus Environment

Page 63 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Au/γ-Al2O3 catalyst in aerobic oxidation of α,ω-diols to lactones. Appl. Catal. B: Environ. 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 γ-butyrolactone. 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.; Schröder, 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

63

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 64 of 95

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 biomass-derived 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 5-hydroxymethylfurfural in water at ambient

64

ACS Paragon Plus Environment

Page 65 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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-Bos 1910, 29, 403-406. (239) Middendorp, J. A. Sur l'oxyméthylfurfurol. Recl. Trav. Chim. Pays-Bos 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.

65

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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.

66

ACS Paragon Plus Environment

Page 66 of 95

Page 67 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

(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 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 biomass-derived 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. Renew. Sustain. 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. Pro. 2015, 34, 4209-4213. (257) Baradii, A. A.; Kokoh, K. B.; Huser, H.; Lamy, C.; LeÂger, 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 2,5-furandicarboxylate)/cloisite 30B composites.

67

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Korean J. Chem. Eng. 2016, 33, 3267-3272. (259) Dijkman, W. P.; Groothuis, D. E.; Fraaije, M. W. Enzyme-catalyzed 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-2,5-dicarbaldehyde 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. Y.; Hwang, S.; Jegal, J.; Kim, Y. H.; Lee, Y. S. Highly selective catalytic hydrogenation and etherification of 5-hydroxymethyl-2-furaldehyde to 2,5-bis(alkoxymethyl)furans for potential biodiesel production. Synlett 2017, 28, 1-4. (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 bio-derived 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 5-hydroxymethylfurfural to potential

68

ACS Paragon Plus Environment

Page 68 of 95

Page 69 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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 5-ethoxymethylfurfural with acid-base bifunctional hybrid nanospheres. Energy Convers. Manage. 2014, 88, 1245-1251.

69

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 70 of 95

(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 5-ethoxymethylfurfural 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.; Mäki-Arvela, 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. Biomass-derived 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 5-hydroxymethylfurfural 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. Poly. Chem. 2015, 6,

70

ACS Paragon Plus Environment

Page 71 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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 2,5-furandicarboxylic 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 single-crystalline Pd nanocrystals for aerobic oxidation of 5-hydroxymethyl-2-furfural. 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

71

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 72 of 95

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. (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, 1202-1211. (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. Poly. Chem. 2017, 8, 3746-3754. (292) Zeng, C.; Seino, H.; Ren, J.; Hatanaka, K.; Yoshie, N. Bio-based furan polymers with self-healing ability. Macromolecules 2013, 46, 1794-1802. (293) Zeng, C.; Seino, H.; Ren, J.; Hatanaka, K.; Yoshie, N. Self-healing bio-based furan polymers cross-linked with various bis-maleimides. Polymer 2013, 54, 5351-5357. (294)

Ikezaki,

T.;

Matsuoka,

poly(2,5-furandimethylene

R.;

Hatanaka,

K.;

succinate-co-butylene

Yoshie,

succinate)

N.

Biobased

crosslinked by

reversible Diels-Alder reaction. J. Polym. Sci. Part A: Polym. Chem. 2014, 52,

72

ACS Paragon Plus Environment

Page 73 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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:

poly(omega-pentadecalactone-co-butylene-co-carbonate).

Synthesis

of

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.

73

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(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.; Doherty, G. A. O. Syntheses of D- and L-mannose, gulose, and talose via diastereoselective and enantioselective dihydroxylation reactions. J. Org. Chem. 1999, 64, 2892-2893. (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.

74

ACS Paragon Plus Environment

Page 74 of 95

Page 75 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Commun. 1992, 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-(r-hydroxyalkyl)furans:

2-hydroxy-exo-brevicomin. Org. Lett. 2011, 13, 1166-1169.

75

ACS Paragon Plus Environment

Synthesis

of

ACS Catalysis

OH O HO

OH O OH

OH

HO O

O

O

OH

HO HO

n

O H OH

OH O

OH OH

O

Cellulose

OH O

O OH HO Starch

O HO

OH OH H Sucrose

OH

OH O n

Hydrolysis

Hydrolysis

Hydrolysis

OH

O

O HO

OH

Hydrolysis

Hydrolysis

O

H

HO H Fructose

COOH OHC

OH OH

CHO

DFF

COOH

O

FFCA

O

ida Ox

O

OHC

O

O

O HO

OH

O O

O

HO HO

OH

tion

FDCA

O

HO

OH OH

Dehydration

O

H

HO

OH OH

HOOC

OH

Hyd rog ena tion

HO

OH

Isomerization

HO

HO HO

OHOH Glucose

Cellobiose

O

O

HO HO

OH

Maltose

MA

OH

OH

OH

O

O

DHMF

DHMTHF

O

O

DMF

DMTHF

HMF O

O O

H3CO

OCH3

O

O

OH OH

HHMFO

FDMC

O

OH O

NHR'

OH O

O OH BHMF

O

Other reactions

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 76 of 95

OH HO HHD

O R

OH

HO HTO

O

O O

O

O O

OH

O O OBMF

AOOMF

AAMFM

O

Figure 1. Catalytic conversion of biomass-derived HMF into various valuable products. Note: DFF, FDCA, FFCA, MA, FDMC, HHMFO, DMF, DMTHF, DHMTHF, HHD, HTO, BHMF, AAMFM,

AOOMF

2,5-furandicarboxylic

and

OBMF

acid,

are

representative

5-formyl-2-furancarboxylic

2,5-furandimethylcarboxylate,

of

acid,

2,5-diformylfuran, maleic

anhydride,

5-hydroxy-5-(hydroxymethyl)furan-2(5H)-one,

2,5-dimethylfuran, 2,5-dimethyltetrahydrofuran, 2,5-dihydroxymethyltetrahydrofuran, 1-hydroxyhexane-2,5-dione,

1,2,6-hexanetriol,

5-arylaminomethyl-2-furanmethanol,

5,5-bis(hydroxymethyl)furoin,

5-alkanoyloxymethylfurfural

5,5-oxy-(bismethylene)-2-furaldehyde, respectively.

76

ACS Paragon Plus Environment

and

Page 77 of 95

O O MF

OH

OH

OH

O

O OH

O O

DHMTHF

MFA

HMF Selective

OH OH

HO HTO

hydrogenation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

OH

O DMF OH

O O OH

DHMF

HO HDO

DMTHF

O HO HHD

O

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

Figure 3. Possible intermediates in the selective hydrogenation of HMF into DHMF over

77

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 78 of 95

Ru-based Shvo’s catalyst.138 Adapted with permission from reference 138. Copyright 2014 Royal Chemistry Society.

Figure 4. Relationships between the conversion of HMF and the δ values of solvents over Pt/MCM-41.144 Adapted with permission from reference 144. Copyright 2014 Royal Chemistry Society. H OH

H

O O

H

N

Ph

Ir Cp* N Ts

CO2

Ph

HMF

OH

OH O

H HN

Ir Cp* N Ts

DHMF Ph

HCOO H

Ph

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

78

ACS Paragon Plus Environment

Page 79 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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

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

79

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 80 of 95

reference 216. Copyright 2014 Elsevier. OH

O

OH

O HMF H+

Electrocatalytic hydrogenation H2O or H+ , standard potential = 0.12 V

+ e-

DHMF

H

H2 O + e -

H + OH -

HMF + 2H

DHMF

HMF + 2H +

OH O

+ 2e -

HMF + 2H 2O + 2e

DHMF -

DHMF + 2OH -

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

Figure 9. Comparison of the onset potentials of the product formation and the HER on transition metals (a) and post-transition metals (b) in neutral and acidic solutions. Note: the onset potentials for the HER are obtained at a current density of −0.5 mA·cm−2.228 Adapted with permission from reference 228. Copyright 2015 John Wiley and Sons.

80

ACS Paragon Plus Environment

Page 81 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

HO

H O

HO

C O

H O

C OH

Ag

Ag H+

e

HO

H O HMF

HO

e , H+

C O

H+

e

H O

Ag

HO

e , H+

C OH

C OH H

Ag

e , H+

H O DHMF

e , H+ HO

H O

HO

C O H

H O

Ag

OH

C

H Ag

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

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

81

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 82 of 95

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.231 Adapted with permission from reference 231. Copyright 2016 Royal Chemistry Society. OH

O

OH

O

OH

OH

O

O

NaOH or NaH

O

OH

+

Cannizzaro reaction

HMF

DHMF

HMFA

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

OH

H

O O

OH

O

2

O

O

HMF

OH

+ DHMF

Al

O O

OH

O

O

O

O

DFF

Reverse MPV

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

O O

OH Escherichia coli CCZU-K14

OH O

Buffer, Glucose, 30 °C

HMF

DHMF

Figure 15. Selective hydrogenation of HMF into DHMF via a biocatalytic pathway in the 82

ACS Paragon Plus Environment

Page 83 of 95

Selective

hydrogenation

Etherification

presence of Escherichia coli CCZU-K14.254

Etherification

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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

83

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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

84

ACS Paragon Plus Environment

Page 84 of 95

Page 85 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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

HO

OH

DHMF O

O

O x O Diacid ethyl ester

O O

CALB

O

O O

x

n

Polymerization

x=2, Succinate; x=3, Glutarate; x=4, Adipate x=6, Suberate; x=8, Sebacate; x=10, Dodecanedioate

85

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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

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

86

ACS Paragon Plus Environment

Page 86 of 95

Page 87 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

ACS Catalysis

Table 1. Conventional hydrogenation of HMF into DHMF by using H2 as a hydrogen donor Pressure Catalyst

Temperature

Solvent

HMF conversion

DHMF yield

(%)

(%)

Time (h) (bar)

(°C)

Reference

Shvo’s

Toluene

10

90

1

99

99

138

Ru/C

Water

50

60

2/3

100

100

139

Ru/ZrO2

1-Butanol

15

120

6

99

99

140

Ru/Al2O3

1-Butanol/water

27

130

2

92

74.5

141

Ru/CeOx

1-Butanol/water

27

130

2

100

81

141

Ru/ZrO2-MgO

1-Butanol/water

27

130

2

99

93.1

141

Ru/ZrO2-SiO2

Water

5

25

4

98.1

90.4

142

Pt/C

Ethanol

14

23

18



82

112

Pt/Al2O3

Ethanol

14

23

18



85

112

Pt/MCM-41

Water

8

35

2

100

98.9

144

Pd/C

Tetrahydrofuran/water

100

80

20

97

82

145

Ir/TiO2

Water

60

50

3

99

94.9

146

Au/FeOx

Water

30

80

2

96

96

147

87

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Page 88 of 95

Au/Al2O3

Water

65

120

2

100

96

148

Cu/C

Ethanol

50

180

8

70

53.8

152

Cu/SiO2

Methanol

25

100

8

100

97

113

Cu/PMO

Ethanol

50

100

3

100

99

149

Cu/ZnO

1,4-Dioxane

15

100

2

100

99.1

150

RANEYCu

Water

90

90

8

94

86.5

151

RANEYNi

Water

90

90

8

100

60

151

NiFe/CNT

1-Butanol

30

110

18

100

96.1

153

CuNi/Al2O3

Tetrahydrofuran

30

130

6

70.6

62.4

154

CuZn/C

Ethanol

50

180

8

63.9

52.2

152

CuZn

Ethanol

70

120

3

100

95

155

CoAl

Methanol

40

120

4

89.4

83

156

PtCo/HCS

1-Butanol

10

120

2

100

70

157

IrRe/SiO2

Water

8

30

6

99

99

158

PtSn/SnO2/RGO

Ethanol

20

70

1/2

99

99

159

Table 2. Transfer hydrogenation of HMF into DHMF by using formic acid and alcohols as hydrogen donors

88

ACS Paragon Plus Environment

Page 89 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

ACS Catalysis

Hydrogen

Temperature Catalyst

Solvent

HMF conversion

DHMF yield

(%)

(%)

Time (h)

donor

(°C)

Reference

Formic acid

Cm∗Ru(HTsDPEN)

Tetrahydrofuran

40

2

100

99

199

Formic acid

Cp∗Ir(HTsDPEN)

Tetrahydrofuran

40

2

100

99

199

Formic acid

Cp∗Ir(HTsDACH)

Tetrahydrofuran

40

1

100

99

199

Formic acid

Cp∗Ir(NHCPh2C6H4)

Tetrahydrofuran

40

1

100

99

199

Formic acid

Pd/C

Tetrahydrofuran

70

4



94

200

Ethanol

ZrO(OH)2

Ethanol

150

5/2

94.1

83.7

203

Isopropanol

Zr-ATPN

Isopropanol

140

2

99

98

204

Isopropanol

Zr-BPPN

Isopropanol

120

2

99

93

205

Isopropanol

Zr-FDCN

Isopropanol

140

8

100

87

206

Isopropanol

Fe2O3@HAP

Isopropanol

180

10

78.2

72

207

Isopropanol

Co3O4@MC

Isopropanol

140

12

100

97

208

89

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Page 90 of 95

Isopropanol

Au/SiC

Isopropanol

20

4

90

83.7

209

Isopropanol

Ru/Co3O4

Isopropanol

190

6

100

82.8

210

Isopropanol

Ru/ZnAlZr-LDH

Isopropanol

200

1/2

100

94

163

2-Butanol

MZH(Zr/Fe=2)

2-Butanol

150

5

98.4

89.6

215

Methanol

MgO

Methanol

160

3

100

100

216

1,4-Butanediol

Cu/AlOX

1,4-Butanediol

220

1/100

94

93

137

Benzyl alcohol

RuCo/C

Benzyl alcohol

150

10

90.7

86.9

171

90

ACS Paragon Plus Environment

Page 91 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Table 3. Reduction potentials (∆Hf°) of various alcohols

a

Alcohol

∆Hf° (kJ/mol)

Methanol

130.1a

Ethanol

85.4a

Propanol

87.3b

Isopropanol

70.0b

1-Butanol

79.7b

2-Butanol

69.3b

The numerical values of ∆Hf° are calculated according to the definition of ∆Hf°. b The

numerical values of ∆Hf° are obtained from van der Waal et al.213

91

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Page 92 of 95

Table 4. Etherification of DHMF into BAMF in various alcohols Alcohol

Catalyst

Temperature (°C)

Time (h)

DHMF conversion (%)

BAMF yield (%)

Reference

Methanol

HZSM-5 (Si/Al=25)

100

3

100

70

113

Methanol

HZSM-5 (Si/Al=38)

100

3

100

68

113

Methanol

HZSM-5 (Si/Al=300)

100

3

100

69

113

Methanol

HZSM-5 (Si/Al=25)

120

12



68a

113

Methanol

HZSM-5 (Si/Al=25)

140

8



59a

113

Methanol

Amberlyst-15

60

10

99

57

263

Methanol

Amberlyst-15

60

10



50a

263

Ethanol

Amberlyst-15

40

16



80

112

Ethanol

Amberlyst-15

60

5



74

112

Ethanol

Amberlyst-15

60

18



64a

112

92

ACS Paragon Plus Environment

Page 93 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

ACS Catalysis

Ethanol

Amberlyst-15

60

10

99

70

263

Ethanol

Amberlyst-15

60

10



70a

263

Ethanol

Sn-Beta

180

6



68.1a

266

Ethanol

Hf-Beta

120

24



67a

267

Propanol

Amberlyst-15

60

10

99

74

263

Propanol

Amberlyst-15

60

10



72a

263

Propanol

Sn-Beta

180

6



61.3a

266

Isopropanol

Sn-Beta

180

3



82.5a

264

Isopropanol

Zr-Mont

150

1



95a

265

Isopropanol

Sn-Beta

180

3

100

85.6

266

Isopropanol

Sn-Beta

180

6



79.6a

266

1-Butanol

Amberlyst-15

60

10

99

74

263

93

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

a

Page 94 of 95

1-Butanol

Amberlyst-15

60

10



71a

263

1-Butanol

Zr-Mont

150

3



49a

265

1-Butanol

Sn-Beta

180

6



60.5a

266

2-Butanol

Zr-Mont

150

1



96a

265

2-Butanol

Sn-Beta

180

6



73.1a

266

2-Butanol

Hf-Beta

120

1



81a

267

HMF was used as a starting substrate, and it would be firstly converted into DHMF over the appropriate hydrogenation catalysts in the

preparation process of BAMF.

94

ACS Paragon Plus Environment

Page 95 of 95 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Table of contents

The state-of-the-art advancements of various chemocatalytic pathways and biocatalytic pathways for the production and application of biomass-derived 2,5-dihydroxymethylfuran are comprehensively summarized and discussed.

95

ACS Paragon Plus Environment