Next Article in Journal
The Effect of Dust Deposition on the Performance of Photovoltaic Panels
Next Article in Special Issue
Polynomial Regression Model Utilization to Determine Potential Refuse-Derived Fuel (RDF) Calories in Indonesia
Previous Article in Journal
Parameter-Free Model Predictive Current Control for PMSM Based on Current Variation Estimation without Position Sensor
Previous Article in Special Issue
Computational Analysis on Combustion Control of Diesel Engines Fueled with Ester Alcohol Diesel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Selective Hydrogenation of 5-Hydroxymethylfurfural over Heterogeneous Metal Catalysts

1
Institute of Chemical Industry of Forest Products, Chinese Academy of Forestry, Nanjing 210042, China
2
School of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou 225002, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(19), 6793; https://doi.org/10.3390/en16196793
Submission received: 30 June 2023 / Revised: 21 September 2023 / Accepted: 22 September 2023 / Published: 24 September 2023
(This article belongs to the Special Issue Biomass and Biofuel for Renewable Energy)

Abstract

:
Biomass is an excellent renewable organic energy in nature. 5-hydroxymethylfurfural (HMF) is a significant platform chemical derived from biomass. It can be obtained from biomass and has the potential to produce high value-added derivatives. For the past few years, the chemocatalysis pathway has been extensively studied and is the main pathway of HMF transformation. In this paper, the influence factors and reaction mechanisms of different catalyst types on HMF hydrogenation processes were discussed. The latest progress on the efficient catalytic system using hydrogen, alcohol and other hydrogen sources to catalyze HMF was introduced. Future research prospects of catalytic hydrogenation of HMF were also prospected.

1. Introduction

Various environmental developments, such as worldwide temperature rise and the rapid exhaustion of nonrenewable fossil reserves due to the use of fossil resources, oil, natural gas and coal, pose a huge challenge to human development. The search for reproducible resources to reduce dependence on nonrenewable petrification resources is urgent and necessary [1,2,3,4]. Biomass is a plant-based material. Compared with hydro, wind, solar and ocean energy, biomass is the only carbon-based resource in the world that can be continuously regenerated. It has many advantages, such as abundant reserves, relatively low cost, wide application and wide distribution. Biomass can be used to produce a wide range of chemicals, fuels and materials to promote sustainable development [5,6,7]. These advantages have attracted wide attention from academia and industry, especially in response to the growing demand for fuels and chemicals [8,9].
Lignocellulosic biomass is a luxuriant, which is rich in cellulose, hemicellulose and lignin. It can be efficiently converted into bio-based fuels and chemicals [10,11]. 5-Hydroxymethylfurfural (HMF) is produced by the chemical reaction of glucose and fructose. In the United States, it is recognized as one of the important biomass-derived platform compounds [12]. It is a crucial furan-based platform molecule that bridges the gap between petrochemical and biomass resources and is known as the “sleeping giant”. Its physicochemical properties are shown in Table 1 [13,14]. HMF has tremendous synthetic capacity and excellent liveness, which is mainly attributed to its three functional groups: aldehyde group, hydroxymethyl group and furan ring structure containing C=O, C–OH and C=C bonds. These bonds provide diversity for metal-catalyzed hydrogenation reactions and produce a series of important heterocyclic derivatives, including 2,5-dimethylfuran (DMF), 5-hydroxymethylfurfuryl alcohol (MFA), 2,5-dihydroxymethyltetrahydrofuran (BHMTHF) and 2,5-dihydroxymethylfuran (BHMF), which are important high-value chemicals, as shown in Figure 1 [15,16,17,18]. Therefore, the selection of a suitable catalytic system can further improve the selectivity and yield of the conversion to several high-value derivatives.
As a potential way to replace traditional H2-mediated hydrogenation reactions, catalytic transfer hydrogenation (CTH) employs renewable and liquid organic hydrogen donors (e.g., alcohols, formic acid, etc.) so that the experimental operation can refrain from the use of high-pressure hydrogen, which can shorten the complexity and cost of experimental stride, and, thus, has lately attracted enormous attention in biomass conversion, as has been noted in many excellent articles [4,19,20]. Because hydrogenation reactions involve the conversion of hydroxyl, aldehyde and furan groups in the presence of hydrogen molecules, including C=O and C=C hydrogenation, decarbonation, and C-O and C-C hydrogenolysis, HMF hydrogenation provides a number of reaction pathways [15,21,22,23]. Catalytic hydrogenation and hydrolysis reactions have been extensively researched because they can effectively reduce the oxygen content in furans, like other biomass-derived platform molecules. Most of the previous catalyst feedstocks for HMF reduction conversion were based on precious metals, like bright, palladium, platinum and ruthenium, which showed excellent intrinsic hydrogenation activity as catalysts, but they were expensive and lacked sufficient economic advantages because they were limited resources. However non-precious metals like copper, Ni and Fe are more abundant and less expensive, and they are potential metallics [11,24,25,26]. However, the selectivity of certain metals in specific reactions is not ideal, leading to the formation of various by-products. Optimization of non-precious metal-based systems could be achieved by combining with other metals and by selecting appropriate carriers and solvents to make mono-and bimetallic multiphase catalysts. So, as long as their activity, selectivity and stability are very efficacious, they will promote the effective reduction of HMF [9,19,27,28].
Table 1. Physicochemical properties of 5-hydroxymethylfurfural (HMF) [29].
Table 1. Physicochemical properties of 5-hydroxymethylfurfural (HMF) [29].
ChemicalMolecular Formula MolecularWeigh
t/(g·mol−1)
Boiling Point/KMelting
Point/K
Density
/(g·cm−3)
5-hydroxym ethylfurfural
(HMF)
C6H6O3126.11387−389301−3071.24
Therefore, the purpose of this review is to describe the HMF metal catalyst hydrogenation conversion process. Our goal is to explore the diversities in the performance of multiphase catalytic systems with noble metal catalysts and non-precious metal catalysts to improve the selectivity of the given reaction. The specificity of the different active metal groups the role of organic solvents, and the differential role of different organic solvents as hydrogen donors, with bargain consideration of the capacity of non-precious metal-based catalysts, will be reviewed. The research in this paper may provide new ideas for multiphase non-precious metal catalysts for the upgrading reaction of oxygen-containing furans of biological origin using organic solvents as hydrogen donors.

2. Direct Hydrogenation

The essential factor to control the selectivity and yield of HMF reaction products is the selection of the hydrogenation catalyst [6,21,29,30]. The metal active site of the catalyst can effectively facilitate the hydrogenation reaction, and the active hydrogen ions assembled on the metal active site hit the nonsaturable organic groups, and eventually the unsaturated bonds on the catalyst surface are reduced. A number of noble metal (Pt, Ru, Pd, Ag, etc.) and non-precious metal (Cu, Fe, Ni, Co, etc.) mono- and bimetallic multiphase catalysts have matured and have been implemented in HMF hydrogenation [9,31].

2.1. Precious Metals

Pt-based catalysts have won the favor of researchers because of their distinct possibilities and are the primary choice for precious metal catalysts. Ge Gao et al. generated PtNi alloy and immobilized PtNi composition metal on SBA-15 in two steps to synthesize a low PtNi alloy catalyst (PtNi/SBA-15) loaded with SBA-15 (a mesoporous silica molar sieve) [32]. The selective hydrogenation of HMF in PtNi/SBA-15 aqueous impregnant at 303 K and 1.5 MPa H2 for 2 h resulted in a conversion of 77.0%, yielding 68.2% BHMF. The PtNi alloy was studied to generate PtNi by hydrothermal method. PtNi/SBA-15 possesses a sequential mesoporous construction with a tall specific face district, and the PtNi composition metal forms Ptδ-Niδ+ face pairs on SBA-15, which is favorable for hydrogen inactivation and alternative carbonyl hydrogenation. Yawen Shi et al. investigated the catalytic movement and separability Pt1Co-fcc and Pt1Co-hcp catalysts of Pt1Co single-atom composition metal (SAA) with fcc (face-centered cuboid) and hcp (hexagonal close-packed) cobalt crystalline stages for the HDO reaction of HMF, separately, by using density flooding theory (DFT) calculations. The best product on the Pt1Co hcp surface was DHMF, while the best product on the Pt1Co-fcc surface was 2-hexanol, comparing the reactivity and regioselectivity of the HDO reaction for HMF conversion on the Pt1Co-fcc and Pt1Co-hcp surfaces [33]. The work of Yawen Shi’s team was successful in developing HDO catalysts with high performance by precise tuning of crystal phases. By DFT calculations, both H2 dissociation and H migration on the catalyst surface were insignificant, but different crystals of the catalyst behave differently with respect to H. The Pt1Co-fcc surface prefers H2 dissociation, and the Pt1Co-hcp surface was more favorable for the migration of H atoms. The different surface structures, which also exhibited different synergistic effects, also described the role of Pt single atoms in promoting hydrogen dissociation and hydrogenation ability and enhancing the reactivity. This is caused by the unlike nuclear arrangement of the fcc and hcp stages with the difference in the atomic density of the crystalline phase, where Pt unitary atoms exhibit specific synergistic effects with Co on the Pt1Co-fcc surface to accelerate the C1-O1 bond breakage. Jian Wang et al. reported on the BHMF selectivity of Pt catalysts loaded with metal oxides (MgO, ZnO, SiO2, and TiO2) for HMF hydrogenation [21]. The study findings indicate that metals mainly activate H2, while the bases of metal oxides coordinate with HMF’s C-O groups, thereby contributing to the separability of BHMF. Tao Gan et al. reported the nuclear interspersion of Pt in Co nanocrystals (denoted as Pt1/Co), and the SAA catalyst was prepared by the ball-milling method [34]. The catalyst showed excellent catalytic performance for HMF hydrodeoxygenation to DMF, with the percentage of 100 HMF transformation and the percentage of 92.9 DMF separability at 1.0 MPa H2, 180 °C and 2 h. Under the collaborative function of platina and cobalt, the primary pathway in the hydrodeoxygenation process is the hydrolysis of the C-O bond in HMF, resulting in the generation of 2,5-dihydroxymethylfuran. Brenda Ledesma et al. researched the catalytic HMF transformation hydrogenated to prepare DMF on mcm−3 and SBA-15 mesoporous material loaded with bimetallic (PtIr) and monometallic (Pt) catalysts and under mild reaction conditions at 120 °C and 15 atm. The catalytic sites of PtIr alloys are highly positive and alternative for the generation of DMF. The PtIr-CMK-3 catalyst has good activity, selectivity and firmness [24]. Vikanova et al. fabricated serial metal Pt catalysts loaded with various contents of Ce/Zr in a 4/1 Ce/Zr ratio as a carrier, and used ultra-low loading Pt catalyst for selective hydrogenation of HMF to prepare BHMF, which could be produced in a BHMF yield up to 97% in ethanol as a solvent under mild conditions (170 °C, 1 MPaH2) [35]. The addition of the low-loaded metal Pt can form a mutual electronic interaction with the carrier Ce, which greatly enhances the adsorption of hydrogen, resulting in a hydrogen spillover effect and providing an additional adsorption center for the carbonyl group.
The Pd-based catalysts are able to selectively hydrogenate by changing the aldehyde group in HMF. The aldehyde-based hydrogenation liveness of Pd-based catalysts are able to be greatly modified by changing the carrier, adjusting the metal particle size and introducing other metals. An efficient WO3/PdOx− carbon shell composite photocatalyst was synthesized by the solution combustion method (SCM) by Alok Kumar et al. [36]. As shown in Figure 2, the catalyst was served as the hydrogenation HMF’s reaction under obvious light (λ ≥ 420 nm) in H2O/MeOH (3:1) conditions without the use of an external pressurized hydrogen carrier, and the core outcome was BHMF with a yield of 79.2%. The uniform size of the carbon shell (10 nm) was found to favor the firmness of the Pd particles (catalytic active sites) and raise the carrier mobility. Shun Nishimura et al. performed the HMF hydrogenation reaction in tetrahydrofuran/water solvent over a Pd/alumina catalyst employing sodium hypophosphite (Na2PO2) as the hydrogen donor at room temperature [37]. The yields of HMF were 45% and of BHMF were 63% at room temperature (25 °C). Wesley R. Silva et al. obtained a micro/nanostructured material Pd/CSCNT-AC with 85% selectivity for 2,5-dihydroxymethylfuran by combining micron-activated carbon (AC) with cuprate stacked carbon nanotubes (CSCNTs) and Pd0/Pd2+ nanoparticles (Pd) [25]. Pd is more beneficial for the hydrogenation of HMF on the C-O and C-C double bonds and catalyzes the hydrogenation of HMF on the C=O double bond. Jingjing Tan et al. designed a RuPd bimetallic catalyst loaded on reduced graphene oxide to achieve 92.9% yield of entire hydrogenation of HMF to DHMTHF at a temperature of 20 degrees Celsius [16]. Contrapositive with monometallic Ru/RGO or Pd/RGO catalysts, bimetallic RuPd/RGO catalysts showed a significant cryogenic effect in the preparation of DHMTHF by HMF hydrogenation. RuPd alloy is shaped during the catalyst reduction. The high synergistic interaction between ruthenium and platinum results in the formation of Ruδ− − Pdδ+ surface pairs. The electron-rich Ruδ− species play a role in facilitating the adsorption and activation of the aldehyde group (C=O) by contributing one or more electrons to the C=O group. In this literature, the authors synthesized bimetallic RuPd catalysts by solvothermal method. Theoretical calculations showed that the furan ring in the HMF molecule could be made to feed electrons back to the electron-deficient Pdδ+ in the catalyst. The results indicated that the electron transfer cycle between the Ruδ− − Pdδ+ pair and the HMF molecule might be the main reason for the high efficiency exhibited by the catalyst in the hydrogenation of HMF to DHMTHF, i.e., the transfer of electrons from the Ruδ− to the C=O group of HMF and the transfer of electrons from the C=C of HMF furan ring bond to the electronic feedback of Pdδ+. Hong Xie et al. wrote that Pd(OAc)2 catalyzed the HMF’s transition to furfuryl alcohol (FFA), 5-MF and DHMF in 1,4-dioxane solution [38].
Currently, Ru-based catalysts are employed in many studies on the hydrogenation reaction of HMF. Hochan Chang et al. complexed HAH monomers by aldol condensation and performed selective hydrogenation over the Ru catalyst to obtain different functional monomers such as HAH, PHAH and FHAH in high yields (≥91 mol%) [39]. Dinesh Kumar Mishra et al. intended solid mixed metal oxides with cobalt–manganese-based MnCo2O4 spinel as a carrier material for ruthenium nanocatalysts using a simple coprecipitation method [40]. BHMF in the percentage of 98.5 yield and BHMTHF in the percentage of 97.3 yield were prepared by efficiently and selectively catalyzing the formation of HMF using Ru/MnCo2O4 as a single catalyst without the addition of any additives. Kaiyue Ji et al. reported a Ru1Cu single-atom alloy (SAA) catalyst with single ruthenium and atoms on copper nanowires, which displayed an electrochemical reduction capability, converting HMF to DHMF with a remarkable yield of 85.6% [41]. The monatomic ruthenium facilitates the segregation of water to generate H* species as a result, which effectively reacts with HMF through the ECH mechanism to generate DHMF. Li Feng et al. researched the loading of Ru metal synthesis catalysts on different carriers (ZSM-5, y-zeolite, b-zeolite, COK-12, filamentous zeolite, zirconia) for 5-HMF hydrogenation activity (in Figure 3) [42]. The Ru/ZSM-5 catalyst indicate the largest HMF transformation rate (98%) and the highest DMF selectivity (97%) and the reaction was conducted at 180 °C using an ethanol solvent system under a hydrogen pressure of 250 psi for 3 h. Yinghao Wang et al. recorded a Ru-Ir/C alloy catalyst that efficiently hydrogenates HMF to DMF without any additional agent under H2 conditions [43]. Up to 100% HMF conversion and 99% DMF selectivity can be achieved using Ru-Ir alloy nanoparticles (~2.2 nm) dispersed with activated carbon as catalysts at 120 °C. This catalyst has the characteristics of selectivity and reusability. Elise B. Gilcher et al. studied two partial hydrogenation intermediates, I-PHAH-O and ipah-oh, specified and contained into a reaction scheme for the hydrogenation of HAH from HMF over platinum and ruthenium catalysts [23]. This study has increased the understanding of dimer particles in liquid-phase hydrogenation and expanded the understanding of the interaction of different functional groups with platinum and ruthenium catalysts in the presence of water. Elise B. Gilcher et al. showed the synthesis of high-purity (>99%) hydroxymethylfurfural-acetone-hydroxymethylfurfural (HAH) by way of a base-catalyzed HMF hydrogenation reaction over platinum, ruthenium and copper-based catalysts at a temperature range from 313 to 393 K condensation reactions [44].
A number of teams have started to utilize Ag-based catalysts for HMF-catalyzed hydrogenation. Huimin Li et al. used Cu as a medium element to modulate the dispersion and electronic configuration of Ag electrocatalysts [45]. AgCu nanoalloys with PA as the carrier were too complex to achieve tall selective ECH of HMF to DHMF. Among them, the ACu1/PA catalyst showed a selectivity of 94% for DHMF in a wide potential range of 0.261~1.122 V in 0.05 M borate dashpot. Xiaotong H et al. clarified the arrangement of BHMF by electrocatalytic hydrogenation of HMF under a mild situation employing carbon-loaded silver nanoparticles (Ag/C) as a cathode catalyst at 85% yield [46]. The sensitivity of the cathodic potential between the silver-catalyzed hydrogenation of HMF to produce BHMF and the hydrogen–dimerization and hydrogen–precipitation reactions of HMF were investigated. The research of Sonali Panigrahy et al. is based on water as a hydrogen source and carbon-supported nano-silver aerogel (Ag-Ag-CNx) as a metal catalyst to detect the conversion of HMF to HD by cyclization reaction at room temperature and pressure. In this reaction, the conversion of HMF was 78% and HD selectivity was 77% [18]. The silver nanoparticles were interconnected on the CNx sheet to form a leachy mesh structure and the reduction of HMF, obtaining sufficient active sites due to the high porosity of the catalyst.

2.2. Non-Precious Metals

As mentioned previously, noble metal catalysts have shown remarkable property in the selective hydrogenation of biomass-derived HMF [7,16]. However, the high cost of precious metal catalysts has limited their large-scale industrial adhibition to some extent [4,36,47]. Therefore, the exploitation of green and competent non-noble metal catalysts is necessary for HMF-catalyzed hydrogenation reactions. Non-noble metal catalysts, such as copper, iron, cobalt and nickel, are popular because of their low cost and abundant reserves [6,48]. Nickel is considered an ideal transition metal with sufficient hydrogenation activity. The introduction of another promoter metal into the nickel catalyst has proven to be an effective strategy that can greatly enhance the catalytic activity and improve the selectivity of the outcomes. Fanhao Kong et al. developed an approach for the fabrication of carbon-coupled sulfur-modulated Ni metal catalysts by a two-step hydrothermal reaction and carbon-thermal reduction using lignosulfonate as carbon and sulfur precursors [49]. Surface Ni hydroxide and carbon shells contribute to increased electroactivity, achieving up to 96% Faraday efficiency and 96% FDCA yield. Brett Pomeroy et al. made activated carbon loaded nickel-based catalysts with the addition of lanthanum and niobium to the Ni catalyst to boost catalytic activity [28]. Good yields are achieved in the reaction temperature range (170–230 °C), hydrogen pressure 5 MPa, and tetrahydrofuran solvent. Tamara Jurado-Vázquez et al. used well-defined homogeneous Ni catalyst precursors [dippeNi(COD)] with formic acid as a hy-hydrogen shift agent to generate the homologous alcohols in positive yields (≥99%) under pacific criteria (5 h, 120 °C) [50]. Zhaonan Zhang et al. produced serial Ni-based catalysts loaded on hydrothermal carbon spheres and adjusted the Ni-based catalyst activity by regulating the nickel nanoparticle dimension and Ni/NiO proportion, and found the percentage of 94 selectivity of BHMF and the percentage of 93.6 yield of DMF on 10%-Ni/HC catalysts with a larger nickel granular diameter [17]. The interaction between Ni (0) and NiO in the calcined Ni/HC catalyst contributes to the activation and hydrogenation of H2, while NiO contributes to the hydrogenation of C-O bond. The excellent selectivity (90%) of BHMF demonstrates that the carbon carrier inhibits the reactivity of Ni to C=C hydrogenation to some degree and exhibits higher selectivity for C=O hydrogenation without restricting the intrinsic activity of nickel. In this article, carbon materials prepared by sucrose hydrothermal method were selected as the carrier of nickel-based catalysts. The strong synergistic effect between Ni and HC materials, as well as the hydrothermal carbon materials possessing abundant oxygen-containing functional groups, makes the catalyst more favorable for HMF chemisorption. The metal Ni induces the activation of H2, while the Lewis acid site in the catalyst facilitates the chemisorption and activation of C-O or C=O bonds in HMF. Yueyue Li et al. clarified the synthesis of a Ni2In-acid-base bifunctional Ni-based catalyst for the hydrogenation and hydrolysis of HMF to DMF with a DMF yield of 93.2% using Ni base-like double hydroxides (LDHs) as the material (Figure 4) [51]. The Ni2In-acid-base site synergistic catalytic system furnishes a new method for the reasonable project of competent hydrogenation–hydrogenolysis catalysts. Zhanghui Xia et al. prepared a ternary Ni-Al/CoOx−1 catalyst to obtain 96% DMF yield by in situ hydrogenation and deoxygenation of HMF under mild reaction conditions with 2-propanol as the hydrogen provenience [52]. Ning Ma et al. successfully prepared multifunctional NiCoTi metal oxide catalysts by heat treatment of NiCoTi layered double hydroxide (LDH) precursors [53]. The hydrogenation transformation of HMF was the percentage of 90.7 and the selectivity of DMF was 95.8% at 200 °C, 1.5 MPa, and the reaction time was 6 h, which showed excellent catalytic performance. María V. Morales et al. prepared four different nickel-based catalysts by using carbonaceous carriers with diverse graphite and porous structures, compared them in HMF selective hydrogenation experiments and found that there is a large impact of the carbon stuff on the catalyst function [54]. The experiments on the possible scenarios showed that the reaction effect of Ni for C=C hydrogenation is to some extent determined by the carbon material, which can lead to high selectivity for carbonyl hydrogenated compounds (DHMF) and inhibition for hydrogenated derivatives (DHMTHF). Finally, the authors also found that metal–carrier interactions display a great effect in modulating the C=O/C=C hydrogenation reaction, demonstrating that different carriers have diverse outcomes on the catalyst performance, and that the selection of a good carrier can help the experimental process.
Cu is attractive for HMF selective hydrogenation because it possesses an entire quantivalence D waveband, it is an oxyphilic metal catalyst that can exert a strong repulsive strength with the carbonyl carbon atom, making it better than C=O hydrogenation and less influential to cycloaddition and decarbonization formation. Consequently, cu-based catalysts have demonstrated excellent performance in the selective transfer hydrogenation of 5-HMF. Xin Yuan et al. conducted a combined experimental and computational study employing Cu electrodes to clarify the crux mechanistic diversity between electrochemical hydrogenation and hydrogenolysis of hydroxymethyl furfural [55]. The analysis of pH effects on the conversion of HMF to various reduction products led to an improvement in the Faraday efficiency of DMF production, reaching 28%. Tong Wang et al. explored the catalytic transfer hydrogenation of 5-HMF in the liquid phase for the delivery of DMF employing isopropanol as the hydrogen donor over a ternary CuxZnAl catalyst [56]. At 180 °C and 4 h of reaction, the DMF yield reached 91.7%. The synergistic effect of surface reduced Cu species (Cu0 and Cu+ species) in the lessened samples was in charge of the strong hydrogenation/hydrogenolysis activity of 5-HMF, especially the positive correlation between electrophilic Cu+ species and DMF yield. Zihao Zhang et al. exploited a ternary CuZnCoOx catalyst with 99% DMF yield for HMF’s in situ hydrogenation employing ethanol as an economic hydrogen donor [57]. The coexistence of the copper and cobalt categories obviously improved the original position of hydrogenation activity of HMF, and the DMF yield reached 99.0% after 5 h of reaction at 210 °C. Kasanneni Tirumala Venkateswara Rao et al. investigated the synthesis of various Cu-alumina (Cu-Al2O3) catalysts by a single procedure of solid-state sharp of Cu and alumina predecessor with oxalic acid, followed by burning and reduction [14]. Under the optimized reaction conditions (3 MPa H2, 130 °C, 1 h), the transformation rate of HMF was 99% and the BHMF yield was 93% using the 20CA (20 mol% Cu-Al2O3) catalyst. The existence of electrophilic Cu(Cu0/Cu2+) and homogeneously sporadic Cu nanoparticles demonstrated that the coexistence of the Cu0/Cu2+ category in the study was responsible for their high selectivity in producing BHMF activity. The potential reaction mechanism for the selective hydrogenation of 5-hydroxymethylfurfural over the copper–alumina catalyst is shown in Figure 5. Jinsung Kim et al. studied the efficiency of the mesoporous Cu-Al2O3 catalyst (me-so-CuA) fixed by the solvent-loss precipitation method in selective hydrogenation of HMF to BHMF [19]. The BHMF’s yield was maintained above 90% for 100 h at 100 °C, 50 bar H2 and 0.2 h of WHSV. Laura M. Esteves et al. prepared 2,5-dimethylfuran (DMF) over different copper-loaded catalysts [58]. A series of Cu-based catalysts Cu/Al2O3, Cu/Fe2O3-Al2O3 and Cu/Nb2O5-Al2O3-623 were fixed, and the HMF was completely converted after 1 h of reaction at 423 K and 20 bar H2 with DMF yields higher than 85%. It was confirmed that a high copper distribution and Lewis acid stand are essential for the stimulation of hydroxyl oxygen in HMF and BHMF to promote DMF formation, and the nature of proper metal dispersion and acid sites display a crucial effect in catalytic activity, providing researchers a better understanding of the function of Cu-based metal catalysts.
Co can be used as the active hydrogenation site. Jing Xia et al. prepared a novel compound metal oxide catalyst (CuCoNiAl-MMO) that involved Co by roasting a decked double hydroxide (LDH) precursor at 500 °C under N2 conditions and applied it to the hydrogenolysis reaction of HMF for the preparation of DMF [59]. CuCoNiAl-MMO exhibited excellent initial activity and selectivity in the hydrogenolysis of HMF, achieving a remarkable 99.8% HMF conversion and 95.3% DMF selectivity at 180 °C, 1.0 MPa H2, and 6.0 h. Shuang Xiang et al. clarified a unique core–shell catalyst, Co@CoO, with the ability to facilitate both homolytic- and heterolytic-integrated cleavage of H2. This catalyst exhibits remarkable capability for the hydrogenolysis of HMF to DMF under pacific cases [60]. Zhao Wenguang et al. successfully prepared bimetallic nanoparticles employing single nitrogen-doped carbon catalysts. They used calcined Cu-Co-zif-9 to immobilize serial Co and Cu nanoparticles and realized in situ formation of metal Co and Cu on nitrogen-doped carbon substrates, forming a Cu-Co/N-C structure with adjustable Co content. The transformation of HMF was 93.7% and the BHMF selectivity was 92.4% at an inferior reaction temperature and H2 pressure. Co nanoparticles are highly dispersed in the N-C skeleton, Co-nx species are formed, and the synergistic catalysis between Cu-Co sites, generated by electron transfer, helps to improve BHMF selectivity [12]. Zhao’s team designed a nitrogen-doped carbon bimetallic catalyst, which opens a new way to realize the efficient conversion and utilization of biomass resources. This bimetallic catalyst prepared with a cubic Cu-Co core and porous structure, a high dispersion of Co nanoparticles in the N-C framework, the formation of Co-nx species, and the synergistic catalytic effect of Cu and Co are favorable for the hydrogenation of HMF into BHMF. In the article, not only the parameters such as the reaction time were optimized but also the effect of the calcination temperature was explored, and the high calcination temperature led to the Co nano-nano particles (NPs) agglomeration, which reduces the catalytic activity of HMF hydrogenation. Jiashi Wang et al. prepared nitrogen-doped graphene shell telescoped metal Co (Co@NGs) catalysts by carbonization and used them for the HMF’s hydrogenation to prepare DMF biofuel. The nitrogen content on the carbon base is efficiently regulated without changing the configuration of Co@NGs or even the metal Co content [61]. Strong electronic interactions are exhibited at the boundary between the outside metal cobalt and the admixed nitrogen. Nearly 100% HMF conversion and 94.7% DMF selectivity were observed at 4 h and 2 MPa hydrogen pressure, which proved that the best Co @ NGs catalyst exhibited high catalytic activity and selectivity. Rosine Ahishakiye et al. developed a fresh Co-CoOx-FeNiCo/γ-Al2O3 catalyst for the selective hydrogenation of DMF from HMF [13]. The catalysts lessened at 500 °C showed excellent catalytic performance with 100% HMF conversion and the largest DMF yield (>99.9%) under the condition that the temperature is 190 degrees and the time is 4 h. The hydrogenation of C-O bonds and separation of hydrogen molecules in HMF is carried out by metal sites, while the hydrogenation of C-O bonds in BHMF and MFA is carried out by CoOx acid sites. Lu Wang et al. researched a competent and cheap zirconium-based metal-organic backbone-supported Co catalyst that was employed to prepare DHMF by catalytic transfer hydrogenation of biomass-derived HMF [27]. By reacting at 100 °C for 4 h, isopropanol as a hydrogen donor and reaction solvent showed outstanding catalytic transfer hydrogenation activity, in which the conversion of HMF reached 92.6% and the selectivity of DHMF was 95.9%. It can be seen that the highly sporadic Co can effectively regulate the acidity and alkalinity of the catalyst, so that the catalyst has higher hydrogenation activity.
Many researchers have started to work on Fe-based catalysts because of its low price and efficient catalytic performance. Jing Xia et al. synthesized a CoNCx/NiFeO catalyst by sorptiving cobalt(II) phthalocyanine (CoPc) on ultrathin nfe-layered double hydroxide nanosheets (nfe–ldh), and then the consequent CoPc/nfe–ldh composites were gently pyrolyzed at 550 °C for 5 h under N2 [20]. CoNCx/NiFeO exhibited excellent performance in the hydrogenation of HMF in tetrahydrofuran (THF) at 180 °C, 1.0 MPa H2, and 6.0 h, achieving a remarkable 99.8% HMF transformation with a selectivity of 94.3% towards DMF. Naimeng Chen et al. formed gold pyroxene (Ni-Fe alloy) on the face of BN by the deposition–precipitation method, while gold augite and Co were consistently distributed on the appearance of BN to compound polymetallic nanoparticle catalysts [62]. It can effectively catalyze the synthesis of DMF from HDO of HMF at 180 °Cwith a yield of 94.0%.
In general, although noble metal catalysts still occupy half of the HMF catalytic hydrogenation, the role of non-precious metal catalysts in the hydrogenation of HMF is becoming more and more virtual. Today, the electronic and spatial effects are regulated by combining precious metals with non-precious metals and a variety of non-precious additives to enhance the divergence and hydrogenation activity of the active metal components. The firmness and dispersion of metal nanoparticles are also improved by changing the metal carrier to strengthen the mass transfer efficiency of reactants and products.

3. Catalytic Transfer Hydrogenation (CTH)

Although there are many advantages of using high pressure H2 as a provenience of hydrogen, HMF has a high conversion rate and excellent results. Many HMF catalytic hydrogenation will use H2 as the hydrogen source, as shown in the Table 2. But the same disadvantages are also evident in dangerous and expensive H2, prompting investigators to seek a cheap and secure selective inventory of hydrogen [1,29]. In recent years, regarding the CTH reaction, hydrogen donors, for instance alcohol, formic acid and silane, have matured and been implemented in organic synthesis one after another. Therefore, the development of various hydrogen donors has become the focus of research in the catalytic conversion of HMF for the replacement of H2 [7,23,54]. In this section, the mechanism of HMF hydrogenation reaction with different hydrogen donors is reviewed and discussed.

3.1. Hydrogen Donor as Alcohol

Low-order alcohols offer crucial advantages as a common hydrogen donor, such as low cost, abundance, ease of repository, and ease of movement. In addition, alcohols also act as solvents in the reaction, facilitating the CTH process and contributing significantly to the catalytic activity of HMF. Yuewen Shao et al. synthesized cu-based nanocatalysts based on LDH to hydrogenate HMF to 1,2-hdo at 120~180 °C in the reaction medium of isopropanol, H2O, and tetrahydrofuran [22]. Shenghui Zhou et al. reported the synthesis of succession acid-based bifunctional metal–lignosulfonate hybrids under hydrothermal conditions and their application to the CTH reaction of 5-hydroxymethylfurfural with 2-propanol; the reaction mechanism is shown in Figure 6 [63]. Under pacific reaction conditions (100 °C, 2 h), the highest catalytic activity was procured for 5-HMF CTH near to the quantitative yield (90%). Yao-Bing Huang et al. used a skeletal CuZnAl catalyst to activate ethanol and hydrogenated HMF to BHMF [30]. When the reaction temperature is 100~120 °C, the yield of BHMF can reach more than 90%. Yung Wei Hsiao et al. revealed highly dispersed Cu/PBSAC catalysts consisting of small metal Cu0 nanoparticles for isopropyl alcohol (IPA) dehydrogenation and subsequent HDO preliminary by HMF [15]. The intermittent reaction using Cu/PBSAC at 190 °C resulted in 91.9% HMF conversion and 71.7% DMF selectivity within 6 h. Yung Wei Hsiao et al. developed a PBSAC-loaded Cu-based catalyst for fixed-bed continuous treatment using a simple initial wet impregnation method. It was proved by calculations that decreasing the Cu-metal particle size and increasing the carrier surface area were the key factors to achieve high DMF yields, where the small particle size was more active for IPA dehydrogenation and HDO of HMF. Moreover, the analysis of the intermediates also proved the effect of external hydrogenation, where IPA dehydrogenation is the rate-limiting step of the reaction, and external hydrogenation can further increase the reaction rate. The final results showed that IPA is also an excellent hydrogen supply solvent for HMF. Aiyong He et al. compounded a novel zirconium–carbon coordination catalyst (Zr-HTC) using a simple self-assembly method, which showed excellent catalytic activity for the selective synthesis of BHMF in isopropanol (iPrOH) as a hydrogen donor; the reaction mechanism is shown in Figure 7 [64]. After heating at a steady temperature of 120 °C for 4 h, a 99.2% yield of BHMF was obtained. Although alcohols like methanol and ethanol have a number of benefits as hydrogen donors, they still have some drawbacks compared to isopropanol [10,15,29]. Methanol requires higher reaction conditions compared to other alcohols, and ethanol may have competitive adsorption with HMF on the catalyst surface, so isopropanol has been studied more as a hydrogen donor [4,7].

3.2. Hydrogen Donor as Acid

Formic acid is obtained by oxidation of biomass or acid hydrolysis, which has the features of extensive source and easy transportation. Tamara Jurado-Vázquez et al. catalyzed the hydrogenation of HMF to the relevant alcohol by using formate as a hydrogen donor under pacific conditions (5 h, 120 °C) with the catalyst [dippeNi(COD)] (dippe = 1,2-bis(diisopropylphosphino)ethane) in positive yields (≥99%) [50]. Jorge Cortez-Elizalde et al. catalyzed the hydrogenation of HMF to DMF with Al2O3TiO2-ZrO2 (Ni/ATZ) composite oxides as carriers and formic acid as the hydrogen donor under argon (Ar) at 210 °C and 20 bar pressure for 24 h. The best results were procured by catalyzing the hydrogenation of HMF to DMF. Formic acid can be produced from the rehydration of HMF from lignocellulosic biomass and is a new type of research approach [65]. Sudeep Mudhulu et al. systematically reviewed the research progress of formic acid-mediated HMF hydrodeoxygenation processes (especially C-¼O bond hydrogenation and CeO bond hydrogenation) in recent years by reviewing the developed heterogeneous catalysts and reaction conditions at home and abroad [66]. It is shown that formic acid has been widely studied as a potential alternative to hydrogen molecules as a liquid organic hydrogen donor for the transfer hydrogenation of biomass derivatives, which contributes to the study of formic acid as a hydrogen donor [3,4]. However, some of the acids are corrosive, so there are not as many studies on acids as hydrogen donors as there are alcohols [22].

3.3. The Hydrogen Donor Is Other

Silane can be activated by varying catalysts under mild conditions to generate H+ and act as a hydrogen donor. Zhaonan Zhang et al. reduced 10%-Ni/HC catalyst through controlling the Ni nanoparticle size and Ni/NiO rate for the hydrogenation of HMF to BHMF under pacific reaction requirements with the best performance of 88% yield (94% selectivity for BHMF) with silane as the hydrogen donor [17]. The reaction can occur without any solvent in the reactor. Youdi Yang et al. revealed that the Pd/C catalyst formatted from γ-valerolactone (Pd-gvl/C) was an effective catalyst for the hydrolysis of HMF to DMF, and 95.6% DMF was procured at 80 °C without the addition of any additives [67]. An efficient catalyst without any additive under mild conditions is the ideal catalyst for HMF hydrolysis to DMF.

4. Conclusions

In summary, the research on the catalytic hydrogenation reaction of HMF under different catalytic systems has made great progress in recent years. HMF is recognized by the U.S. Department of Energy as a vital flatform chemical with great potential to replace nonrenewable resources in the composition of materials, fuels and crucial chemicals. However, there are still a number of problems in the catalytic conversion process. First, the attention to the hydrogenation reaction of other unsaturated groups (other than carbon, nitrogen and oxygen functional groups) should be enhanced. Subsequently, for the catalytic hydrogenation reaction of HMF, this part of unsaturated functional groups will be the major breakthrough point. Secondly, the combination of density functional theory, molecular dynamics and energetics is significant for studying the reaction mechanism. Finally, the catalyst preparation method and reaction system are the key to HMF conversion. Various efficient non-precious metal catalysts and green, safe and efficient hydrogen donors are under research. In summary, with sustainable development as the goal, the industrial production of HMF-derived fuels and chemicals will make more progress and breakthroughs in the future.

Author Contributions

Conceptualization, H.X.; writing-original draft preparation, H.X. and J.L.; Writing-review andediting, M.Z.; funding acquisition, M.Z. and H.X. All authors have read and agreed to the published version of the manuscript.

Funding

Authors are grateful for the financial support from the National Natural Science Foundation of China (31971609) and the Fundamental Research Funds of CAFYBB2022MA002.

Data Availability Statement

Data is available if necessary.

Conflicts of Interest

There are no conflict of interest.

References

  1. Chen, S.; Wojcieszak, R.; Dumeignil, F.; Marceau, E.; Royer, S. How Catalysts and Experimental Conditions Determine the Selective Hydroconversion of Furfural and 5-Hydroxymethylfurfural. Chem. Rev. 2018, 118, 11023–11117. [Google Scholar] [CrossRef]
  2. Fang, W.; Riisager, A. Recent advances in heterogeneous catalytic transfer hydrogenation/hydrogenolysis for valorization of biomass-derived furanic compounds. Green Chem. 2021, 23, 670–688. [Google Scholar] [CrossRef]
  3. Turkin, A.; Eyley, S.; Preegel, G.; Thielemans, W.; Makshina, E.; Sels, B.F. How Trace Impurities Can Strongly Affect the Hydroconversion of Biobased 5-Hydroxymethylfurfural? ACS Catal. 2021, 11, 9204–9209. [Google Scholar] [CrossRef]
  4. Messori, A.; Fasolini, A.; Mazzoni, R. Advances in Catalytic Routes for the Homogeneous Green Conversion of the Bio-Based Platform 5-Hydroxymethylfurfural. ChemSusChem 2022, 15, e202200228. [Google Scholar] [CrossRef] [PubMed]
  5. Hu, L.; Lin, L.; Wu, Z.; Zhou, S.; Liu, S. Recent advances in catalytic transformation of biomass-derived 5-hydroxymethylfurfural into the innovative fuels and chemicals. Renew. Sustain. Energy Rev. 2017, 74, 230–257. [Google Scholar] [CrossRef]
  6. Wang, X.; Liang, X.; Li, J.; Li, Q. Catalytic hydrogenolysis of biomass-derived 5-hydroxymethylfurfural to biofuel 2, 5-dimethylfuran. Appl. Catal. A Gen. 2019, 576, 85–95. [Google Scholar] [CrossRef]
  7. Weng, R.; Lu, X.; Ji, N.; Fukuoka, A.; Shrotri, A.; Li, X.; Zhang, R.; Zhang, M.; Xiong, J.; Yu, Z. Taming the butterfly effect: Modulating catalyst nanostructures for better selectivity control of the catalytic hydrogenation of biomass-derived furan platform chemicals. Catal. Sci. Technol. 2021, 11, 7785–7806. [Google Scholar] [CrossRef]
  8. Park, D.; Lee, S.; Kim, J.; Ryu, G.Y.; Suh, Y. 5-(Chloromethyl)Furfural as a Potential Source for Continuous Hydrogenation of 5-(Hydroxymethyl)Furfural to 2,5-Bis(Hydroxymethyl)Furan. ChemPlusChem 2022, 87, e202200272. [Google Scholar] [CrossRef]
  9. Huang, Z.; Wang, J.; Lei, J.; Zhao, W.; Chen, H.; Yang, Y.; Xu, Q.; Liu, X. Recent Advances in the Catalytic Hydroconversion of 5-Hydroxymethylfurfural to Valuable Diols. Front. Chem. 2022, 10, 925603. [Google Scholar] [CrossRef] [PubMed]
  10. Kong, X.; Zhu, Y.; Fang, Z.; Kozinski, J.A.; Butler, I.S.; Xu, L.; Song, H.; Wei, X. Catalytic conversion of 5-hydroxymethylfurfural to some value-added derivatives. Green Chem. 2018, 20, 3657–3682. [Google Scholar] [CrossRef]
  11. Hu, L.; He, A.; Liu, X.; Xia, J.; Xu, J.; Zhou, S.; Xu, J. Biocatalytic Transformation of 5-Hydroxymethylfurfural into High-Value Derivatives: Recent Advances and Future Aspects. ACS Sustain. Chem. Eng. 2018, 6, 15915–15935. [Google Scholar] [CrossRef]
  12. Zhao, W.; Zhu, X.; Zeng, Z.; Lei, J.; Huang, Z.; Xu, Q.; Liu, X.; Yang, Y. Cu-Co nanoparticles supported on nitrogen-doped carbon: An efficient catalyst for hydrogenation of 5-hydroxymethylfurfural into 2,5-bis(hydroxymethyl)furan. Mol. Catal. 2022, 524, 112304. [Google Scholar] [CrossRef]
  13. Ahishakiye, R.; Wang, F.; Zhang, X.; Sun, M.; Zhai, Y.; Liu, Y.; Wu, Y.; Li, M.; Li, M.; Zhang, Q. Novel noble metal-free and recyclable Co-CoOx-FeNiCo/γ-Al2O3 catalyst for selective hydrogenation of 5-hydroxymethylfurfural to 2,5-dimethylfuran or 2,5-Bis(hydroxymethyl)furan. Chem. Eng. J. 2022, 450, 138187. [Google Scholar] [CrossRef]
  14. Rao, K.T.V.; Hu, Y.; Yuan, Z.; Zhang, Y.; Xu, C.C. Green synthesis of heterogeneous copper-alumina catalyst for selective hydrogenation of pure and biomass-derived 5-hydroxymethylfurfural to 2,5-bis(hydroxymethyl)furan. Appl. Catal. A Gen. 2021, 609, 117892. [Google Scholar] [CrossRef]
  15. Hsiao, Y.W.; Zong, X.; Zhou, J.; Zheng, W.; Vlachos, D.G. Selective hydrodeoxygenation of 5-hydroxymethylfurfural (HMF) to 2,5-dimethylfuran (DMF) over carbon supported copper catalysts using isopropyl alcohol as a hydrogen donor. Appl. Catal. B Environ. 2022, 317, 121790. [Google Scholar] [CrossRef]
  16. Tan, J.J.; Cui, J.; Zhu, Y.; Cui, X.; Shi, Y.; Yan, W.; Zhao, Y. Complete Aqueous Hydrogenation of 5-Hydroxymethylfurfural at Room Temperature over Bimetallic RuPd/Graphene Catalyst. ACS Sustain. Chem. Eng. 2019, 7, 10670–10678. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Liu, C.; Liu, D.; Shang, Y.; Yin, X.; Zhang, P.; Mamba, B.B.; Kuvarega, A.T.; Gui, J. Hydrothermal carbon-supported Ni catalysts for selective hydrogenation of 5-hydroxymethylfurfural toward tunable products. J. Mater. Sci. 2020, 55, 14179–14196. [Google Scholar] [CrossRef]
  18. Panigrahy, S.; Mishra, R.; Panda, P.; Kempasiddaiah, M.; Barman, S. Carbon-Supported Ag Nanoparticle Aerogel for Electrocatalytic Hydrogenation of 5-(Hydroxymethyl)furfural to 2,5-Hexanedione Under Acidic Conditions. ACS Appl. Nano Mater. 2022, 5, 8314–8323. [Google Scholar] [CrossRef]
  19. Kim, J.; Bathula, H.B.; Yun, S.; Jo, Y.; Lee, S.; Baik, J.H.; Suh, Y.-W. Hydrogenation of 5-hydroxymethylfurfural into 2,5-bis(hydroxymethyl)furan over mesoporous Cu–Al2O3 catalyst: From batch to continuous processing. J. Ind. Eng. Chem. 2021, 102, 186–194. [Google Scholar] [CrossRef]
  20. Xia, J.; Gao, D.; Han, F.; Li, Y.; Waterhouse, G.I.N. Efficient and Selective Hydrogenation of 5-Hydroxymethylfurfural to 2,5-Dimethylfuran Over a Non-noble CoNCx/NiFeO Catalyst. Catal. Lett. 2022, 152, 3400–3413. [Google Scholar] [CrossRef]
  21. Wang, J.; Zhao, J.; Fu, J.; Miao, C.; Jia, S.; Yan, P.; Huang, J. Highly selective hydrogenation of 5-hydroxymethylfurfural to 2,5-bis(hydroxymethyl)furan over metal-oxide supported Pt catalysts: The role of basic sites. Appl. Catal. A Gen. 2022, 643, 118762. [Google Scholar] [CrossRef]
  22. Shao, Y.; Wang, J.; Sun, K.; Gao, G.; Fan, M.; Li, C.; Ming, C.; Zhang, L.; Zhang, S.; Hu, X. Cu-Based Nanoparticles as Catalysts for Selective Hydrogenation of Biomass-Derived 5-Hydroxymethylfurfural to 1,2-Hexanediol. ACS Appl. Nano Mater. 2022, 5, 5882–5894. [Google Scholar] [CrossRef]
  23. Gilcher, E.B.; Chang, H.; Rebarchik, M.; Huber, G.W.; Dumesic, J.A. Effects of Water Addition to Isopropanol for Hydrogenation of Compounds Derived from 5-Hydroxymethyl Furfural over Pd, Ru, and Cu Catalysts. ACS Catal. 2022, 12, 10186–10198. [Google Scholar] [CrossRef]
  24. Ledesma, B.; Juárez, J.; Mazarío, J.; Domine, M.; Beltramone, A. Bimetallic platinum/iridium modified mesoporous catalysts applied in the hydrogenation of HMF. Catal. Today 2021, 360, 147–156. [Google Scholar] [CrossRef]
  25. Silva, W.R.; Matsubara, E.Y.; Rosolen, J.M.; Donate, P.M.; Gunnella, R. Pd catalysts supported on different hydrophilic or hydrophobic carbonaceous substrate for furfural and 5-(hydroxymethyl)-furfural hydrogenation in water. Mol. Catal. 2021, 504, 111496. [Google Scholar] [CrossRef]
  26. Turkin, A.A.; Makshina, E.V.; Sels, B.F. Catalytic Hydroconversion of 5-HMF to Value-Added Chemicals: Insights into the Role of Catalyst Properties and Feedstock Purity. ChemSusChem 2022, 15, e202200412. [Google Scholar] [CrossRef]
  27. Wang, L.; Zuo, J.; Zhang, Q.; Peng, F.; Chen, S.; Liu, Z. Catalytic Transfer Hydrogenation of Biomass-Derived 5-Hydroxymethylfurfural into 2,5-Dihydroxymethylfuran over Co/UiO-66-NH2. Catal. Lett. 2021, 152, 361–371. [Google Scholar] [CrossRef]
  28. Pomeroy, B.; Grilc, M.; Gyergyek, S.; Likozar, B. Catalyst structure-based hydroxymethylfurfural (HMF) hydrogenation mechanisms, activity and selectivity over Ni. Chem. Eng. J. 2021, 412, 127553. [Google Scholar] [CrossRef]
  29. Zhang, J.; Li, D.N.; Yuan, H.R.; Wang, S.R.; Yong, C.H.E.N. Advances on the catalytic hydrogenation of bio-mass-derived furfural and 5-hydroxymethylfurfural. J. Fuel Chem. Technol. 2021, 49, 1752–1766. [Google Scholar] [CrossRef]
  30. Huang, Y.; Zhang, X.; Zhang, J.; Chen, H.; Wang, T.; Lu, Q. Catalytic Transfer Hydrogenation of 5-Hydroxymethylfurfural with Primary Alcohols over Skeletal CuZnAl Catalysts. ChemSusChem 2022, 15, e202200237. [Google Scholar] [CrossRef]
  31. Qu, D.; He, S.; Chen, L.; Ye, Y.; Ge, Q.; Cong, H.; Jiang, N.; Ha, Y. Paired electrocatalysis in 5-hydroxymethylfurfural valorization. Front. Chem. 2022, 10, 1055865. [Google Scholar] [CrossRef]
  32. Gao, G.; Jiang, Z.; Hu, C. Selective Hydrogenation of the Carbonyls in Furfural and 5-Hydroxymethylfurfural Catalyzed by PtNi Alloy Supported on SBA-15 in Aqueous Solution Under Mild Conditions. Front. Chem. 2021, 9, 759512. [Google Scholar] [CrossRef] [PubMed]
  33. Shi, Y.; Ding, H.; Wang, S.; Deng, Y.; Feng, H.; Liang, Y.; Chen, L.; Zhang, X. Crystal Phase Sensitivity of 5-Hydroxymethylfurfural Hydrodeoxygenation over a Pt1Co Single-Atom Alloy Catalyst. J. Phys. Chem. C 2022, 126, 20807–20815. [Google Scholar] [CrossRef]
  34. Gan, T.; Liu, Y.; He, Q.; Zhang, H.; He, X.; Ji, H. Facile Synthesis of Kilogram-Scale Co-Alloyed Pt Single-Atom Catalysts via Ball Milling for Hydrodeoxygenation of 5-Hydroxymethylfurfural. ACS Sustain. Chem. Eng. 2020, 8, 8692–8699. [Google Scholar] [CrossRef]
  35. Vikanova, K.V.; Chernova, M.S.; Redina, E.A.; Kapustin, G.I.; Tkachenko, O.P.; Kustov, L.M. Heterogeneous additive-free highly selective synthesis of 2,5-bis(hydroxymethyl)furan over catalysts with ultra-low Pt content. J. Chem. Technol. Biotechnol. 2021, 96, 2421–2425. [Google Scholar] [CrossRef]
  36. Kumar, A.; Ananthakrishnan, R. Strategic Design of a WO3/PdOx-Carbon Shell Composite Photocatalyst: Visible Light-Mediated Selective Hydrogenation of 5-Hydroxymethylfurfural. ACS Appl. Energy Mater. 2022, 5, 2706–2719. [Google Scholar] [CrossRef]
  37. Nishimura, S.; Shimura, T.; Ebitani, K. Transfer hydrogenation of furaldehydes with sodium phosphinate as a hydrogen source using Pd-supported alumina catalyst. J. Taiwan Inst. Chem. Eng. 2017, 79, 97–102. [Google Scholar] [CrossRef]
  38. Xie, H.; Qi, T.; Lyu, Y.-J.; Zhang, J.-F.; Si, Z.-B.; Liu, L.-J.; Zhu, L.-F.; Yang, H.-Q.; Hu, C.-W. Molecular mechanism comparison of decarbonylation with deoxygenation and hydrogenation of 5-hydroxymethylfurfural catalyzed by palladium acetate. Phys. Chem. Chem. Phys. 2019, 21, 3795–3804. [Google Scholar] [CrossRef] [PubMed]
  39. Chang, H.; Gilcher, E.B.; Huber, G.W.; Dumesic, J.A. Synthesis of performance-advantaged polyurethanes and polyesters from biomass-derived monomers by aldol-condensation of 5-hydroxymethyl furfural and hydrogenation. Green Chem. 2021, 23, 4355–4364. [Google Scholar] [CrossRef]
  40. Mishra, D.K.; Lee, H.J.; Truong, C.C.; Kim, J.; Suh, Y.-W.; Baek, J.; Kim, Y.J. Ru/MnCo2O4 as a catalyst for tunable synthesis of 2,5-bis(hydroxymethyl)furan or 2,5-bis(hydroxymethyl)tetrahydrofuran from hydrogenation of 5-hydroxymethylfurfural. Mol. Catal. 2020, 484, 110722. [Google Scholar] [CrossRef]
  41. Ji, K.; Xu, M.; Xu, S.; Wang, Y.; Ge, R.; Hu, X.; Sun, X.; Duan, H. Electrocatalytic Hydrogenation of 5-Hydroxymethylfurfural Promoted by a Ru1Cu Single-Atom Alloy Catalyst. Angew. Chem. Int. Ed. 2022, 61, e202209849. [Google Scholar] [CrossRef]
  42. Feng, L.; Li, X.; Lin, Y.; Liang, Y.; Chen, Y.; Zhou, W. Catalytic hydrogenation of 5-hydroxymethylfurfural to 2,5-dimethylfuran over Ru based catalyst: Effects of process parameters on conversion and products selectivity. Renew. Energy 2020, 160, 261–268. [Google Scholar] [CrossRef]
  43. Wang, Y.; Wang, Y.; Lu, Y.; Cao, Q.-E.; Fang, W. Efficient hydrogenation of 5-hydroxymethylfurfural using a synergistically bimetallic Ru–Ir/C catalyst. Chem. Commun. 2021, 57, 1742–1745. [Google Scholar] [CrossRef] [PubMed]
  44. Gilcher, E.B.; Chang, H.; Huber, G.W.; Dumesic, J.A. Controlled hydrogenation of a biomass-derived platform chemical formed by aldol-condensation of 5-hydroxymethyl furfural (HMF) and acetone over Ru, Pd, and Cu catalysts. Green Chem. 2022, 24, 2146–2159. [Google Scholar] [CrossRef] [PubMed]
  45. Li, H.; Zhang, T.; Peng, M.; Zhang, Q.; Liu, J.; Zhang, J.; Fu, Y.; Li, W. Highly selective electrocatalytic hydrogenation of 5-hydroxymethylfurfural to 2,5-dihydroxymethylfuran over AgCu nanoalloys. Int. J. Hydrogen Energy 2022, 47, 28904–28914. [Google Scholar] [CrossRef]
  46. Chadderdon, X.H.; Chadderdon, D.J.; Pfennig, T.; Shanks, B.H.; Li, W. Paired electrocatalytic hydrogenation and oxidation of 5-(hydroxymethyl)furfural for efficient production of biomass-derived monomers. Green Chem. 2019, 21, 6210–6219. [Google Scholar] [CrossRef]
  47. Xu, C.; Paone, E.; Rodríguez-Padrón, D.; Luque, R.; Mauriello, F. Recent catalytic routes for the preparation and the upgrading of biomass derived furfural and 5-hydroxymethylfurfural. Chem. Soc. Rev. 2020, 49, 4273–4306. [Google Scholar] [CrossRef]
  48. Jiang, Z.; Wan, W.; Lin, Z.; Xie, J.; Chen, J.G. Understanding the Role of M/Pt(111) (M = Fe, Co, Ni, Cu) Bimetallic Surfaces for Selective Hydrodeoxygenation of Furfural. ACS Catal. 2017, 7, 5758–5765. [Google Scholar] [CrossRef]
  49. Kong, F.; Wang, M. Preparation of Sulfur-Modulated Nickel/Carbon Composites from Lignosulfonate for the Electrocatalytic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid. ACS Appl. Energy Mater. 2021, 4, 1182–1188. [Google Scholar] [CrossRef]
  50. Jurado-Vázquez, T.; Arévalo, A.; García, J.J. Furfural and 5-(hydroxymethyl)furfural valorization using homogeneous Ni(0) and Ni(II) catalysts by transfer hydrogenation. J. Organomet. Chem. 2022, 957, 122162. [Google Scholar] [CrossRef]
  51. Li, Y.; Wang, R.; Huang, B.; Zhang, L.; Ma, X.; Zhang, S.; Zhu, Z.; Lü, H.; Yang, K. Hydrogenation and hydrogenolysis of 5-hydroxymethylfurfural to 2,5-dimethylfuran via synergistic catalysis of Ni2In and acid-base sites. Appl. Surf. Sci. 2022, 604, 154579. [Google Scholar] [CrossRef]
  52. Xia, Z.; Niu, L.; An, Y.; Bian, G.; Li, T.; Bai, G. Ni–Al/CoOx-catalyzed hydrodeoxygenation of 5-hydroxymethylfurfural into 2,5-dimethylfuran at low temperatures without external hydrogen. Green Chem. 2021, 23, 7763–7772. [Google Scholar] [CrossRef]
  53. Ma, N.; Song, Y.; Han, F.; Waterhouse, G.I.N.; Li, Y.; Ai, S. Multifunctional NiCoTi Catalyst Derived from Layered Double Hydroxides for Selective Hydrogenation of 5-Hydroxymethylfurfural to 2,5-Dimethylfuran. Catal. Lett. 2020, 151, 517–525. [Google Scholar] [CrossRef]
  54. Morales, M.V.; Conesa, J.M.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Tunable selectivity of Ni catalysts in the hydrogenation reaction of 5-hydroxymethylfurfural in aqueous media: Role of the carbon supports. Carbon 2021, 182, 265–275. [Google Scholar] [CrossRef]
  55. Yuan, X.; Lee, K.; Bender, M.T.; Schmidt, J.R.; Choi, K. Mechanistic Differences between Electrochemical Hydrogenation and Hydrogenolysis of 5-Hydroxymethylfurfural and Their pH Dependence. ChemSusChem 2022, 15, e202200952. [Google Scholar] [CrossRef]
  56. Wang, T.; Li, F.; Xue, W.; Wang, Y. Highly efficient catalytic transfer hydrogenation of 5-hydroxymethylfurfural to 2,5-dimethylfuran over CuxZnAl catalysts. Asia-Pac. J. Chem. Eng. 2021, 17, e2736. [Google Scholar] [CrossRef]
  57. Zhang, Z.; Yao, S.; Wang, C.; Liu, M.; Zhang, F.; Hu, X.; Chen, H.; Gou, X.; Chen, K.; Zhu, Y.; et al. CuZnCoOx multifunctional catalyst for in situ hydrogenation of 5-hydroxymethylfurfural with ethanol as hydrogen carrier. J. Catal. 2019, 373, 314–321. [Google Scholar] [CrossRef]
  58. Esteves, L.M.; Brijaldo, M.H.; Oliveira, E.G.; Martinez, J.J.; Rojas, H.; Caytuero, A.; Passos, F.B. Effect of support on selective 5-hydroxymethylfurfural hydrogenation towards 2,5-dimethylfuran over copper catalysts. Fuel 2020, 270, 117524. [Google Scholar] [CrossRef]
  59. Xia, J.; Gao, D.; Han, F.; Lv, R.; Waterhouse, G.I.N.; Li, Y. Hydrogenolysis of 5-Hydroxymethylfurfural to 2,5-Dimethylfuran Over a Modified CoAl-Hydrotalcite Catalyst. Front. Chem. 2022, 10, 907649. [Google Scholar] [CrossRef]
  60. Xiang, S.; Dong, L.; Wang, Z.-Q.; Han, X.; Daemen, L.L.; Li, J.; Cheng, Y.; Guo, Y.; Liu, X.; Hu, Y.; et al. A unique Co@CoO catalyst for hydrogenolysis of biomass-derived 5-hydroxymethylfurfural to 2,5-dimethylfuran. Nat. Commun. 2022, 13, 3657. [Google Scholar] [CrossRef]
  61. Wang, J.; Wei, Q.; Ma, Q.; Guo, Z.; Qin, F.; Ismagilov, Z.R.; Shen, W. Constructing Co@N-doped graphene shell catalyst via Mott-Schottky effect for selective hydrogenation of 5-hydroxylmethylfurfural. Appl. Catal. B Environ. 2020, 263, 118339. [Google Scholar] [CrossRef]
  62. Chen, N.; Zhu, Z.; Su, T.; Liao, W.; Deng, C.; Ren, W.; Zhao, Y.; Lü, H. Catalytic hydrogenolysis of hydroxymethylfurfural to highly selective 2,5-dimethylfuran over FeCoNi/h-BN catalyst. Chem. Eng. J. 2020, 381, 122755. [Google Scholar] [CrossRef]
  63. Zhou, S.; Dai, F.; Chen, Y.; Dang, C.; Zhang, C.; Liu, D.; Qi, H. Sustainable hydrothermal self-assembly of hafnium–lignosulfonate nanohybrids for highly efficient reductive upgrading of 5-hydroxymethylfurfural. Green Chem. 2019, 21, 1421–1431. [Google Scholar] [CrossRef]
  64. He, A.; Hu, L.; Zhang, Y.; Jiang, Y.; Wang, X.; Xu, J.; Wu, Z. High-Efficiency Catalytic Transfer Hydrogenation of Biomass-Based 5-Hydroxymethylfurfural to 2,5-Bis(hydroxymethyl)furan over a Zirconium–Carbon Coordination Catalyst. ACS Sustain. Chem. Eng. 2021, 9, 15557–15570. [Google Scholar] [CrossRef]
  65. Cortez-Elizalde, J.; Córdova-Pérez, G.E.; Silahua-Pavón, A.A.; Pérez-Vidal, H.; Cervantes-Uribe, A.; Cordero-García, A.; Arévalo-Pérez, J.C.; Becerril-Altamirano, N.L.; Castillo-Gallegos, N.C.; Lunagómez-Rocha, M.A.; et al. 2,5-Dimethylfuran Production by Catalytic Hydrogenation of 5-Hydroxymethylfurfural Using Ni Supported on Al2O3-TiO2-ZrO2 Prepared by Sol-Gel Method: The Effect of Hydrogen Donors. Molecules 2022, 27, 4187. [Google Scholar] [CrossRef] [PubMed]
  66. Mudhulu, S.; Gong, Z.-J.; Ku, H.-C.; Lu, Y.-H.; Yu, W.-Y. Recent advances in heterogeneous catalytic hydrodeoxygenation of biomass-derived oxygenated furanics mediated by formic acid. Mater. Today Sustain. 2022, 19, 100199. [Google Scholar] [CrossRef]
  67. Yang, Y.; Liu, H.; Li, S.; Chen, C.; Wu, T.; Mei, Q.; Wang, Y.; Chen, B.; Liu, H.; Han, B. Hydrogenolysis of 5-Hydroxymethylfurfural to 2,5-Dimethylfuran under Mild Conditions without Any Additive. ACS Sustain. Chem. Eng. 2019, 7, 5711–5716. [Google Scholar] [CrossRef]
Figure 1. Typical chemical properties and conversion routes for HMF.
Figure 1. Typical chemical properties and conversion routes for HMF.
Energies 16 06793 g001
Figure 2. Potential reaction approaches for hydrogenation of HMF to BHMF over the photocatalyst WO3/PdOx@C [36] (Adapted from references [36]).
Figure 2. Potential reaction approaches for hydrogenation of HMF to BHMF over the photocatalyst WO3/PdOx@C [36] (Adapted from references [36]).
Energies 16 06793 g002
Figure 3. Possible reaction pathways HMF hydrogenation into product [42] (Adapted from references [42]).
Figure 3. Possible reaction pathways HMF hydrogenation into product [42] (Adapted from references [42]).
Energies 16 06793 g003
Figure 4. A proposed hydrogenation–hydrogenolysis mechanism for HMF transformation to DMF over Ni2In/MgO-Al2O3 [51] (Adapted from reference [51]).
Figure 4. A proposed hydrogenation–hydrogenolysis mechanism for HMF transformation to DMF over Ni2In/MgO-Al2O3 [51] (Adapted from reference [51]).
Energies 16 06793 g004
Figure 5. The plausible reaction mechanism proposed for selective hydrogenation of 5-HMF to BHMF over the copper–alumina catalyst [14] (Adapted from reference [14]).
Figure 5. The plausible reaction mechanism proposed for selective hydrogenation of 5-HMF to BHMF over the copper–alumina catalyst [14] (Adapted from reference [14]).
Energies 16 06793 g005
Figure 6. Possible mechanism of the 5-HMF reduction reaction catalyzed by 2-propanol as a hydrogen donor [63] (Adapted from references [63]).
Figure 6. Possible mechanism of the 5-HMF reduction reaction catalyzed by 2-propanol as a hydrogen donor [63] (Adapted from references [63]).
Energies 16 06793 g006
Figure 7. Plausible mechanism for the selective synthesis of BHMF from HMF via CTH in iPrOH over Zr-HTC [64] (Adapted from reference [64]).
Figure 7. Plausible mechanism for the selective synthesis of BHMF from HMF via CTH in iPrOH over Zr-HTC [64] (Adapted from reference [64]).
Energies 16 06793 g007
Table 2. Hydrogenation of HMF using molecular H2 as hydrogen source.
Table 2. Hydrogenation of HMF using molecular H2 as hydrogen source.
EntryCatalystSolventTime/hTemp/KPress/MPaConv/%ProductYield/% Ref
1PtNi/SBA-15H2O23031.577.0%BHMF68.2%[32]
2Pt1/CoTHF24531100%DMF92.9%[34]
3Ru/ZSM-5ethanol34531.7298%DMF97%[42]
4Ru-Ir/CTHF123931100%DNF99%[43]
5Ag−aerogel−CNxH2O3303078%HD77%[18]
6Ni/HHTH2O4333398%BHMF90%[54]
720 mol%
Cu-Al2O3
methanol1403399%BHMF93%[14]
8meso-CuAethanol0.2373590%DMF85%[19]
9CuCoNiAl-MMOTHF6453199.8DMF95.3[59]
10Cu-Co/N-CTHF4373193.7BHMF92.4%[12]
11Co@NGsethanol44732100%DMF94.7%[61]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xia, H.; Li, J.; Zhou, M. Advances in Selective Hydrogenation of 5-Hydroxymethylfurfural over Heterogeneous Metal Catalysts. Energies 2023, 16, 6793. https://doi.org/10.3390/en16196793

AMA Style

Xia H, Li J, Zhou M. Advances in Selective Hydrogenation of 5-Hydroxymethylfurfural over Heterogeneous Metal Catalysts. Energies. 2023; 16(19):6793. https://doi.org/10.3390/en16196793

Chicago/Turabian Style

Xia, Haihong, Jing Li, and Minghao Zhou. 2023. "Advances in Selective Hydrogenation of 5-Hydroxymethylfurfural over Heterogeneous Metal Catalysts" Energies 16, no. 19: 6793. https://doi.org/10.3390/en16196793

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

Article Metrics

Back to TopTop