Kangyu Zhao,
Bin Wen,
Qing Tang,
Feng Wang,
Xianxiang Liu*,
Qiong Xu and
Dulin Yin
National & Local Joint Engineering Laboratory for New Petro-chemical Materials and Fine Utilization of Resources, Key Laboratory of the Assembly and Application of Organic Functional Molecules of Hunan Province, Hunan Normal University, Changsha 410081, China. E-mail: lxx@hunnu.edu.cn
First published on 16th May 2024
To address the problem of non-renewable resources and energy shortages, converting biomass, the only renewable carbon resource on Earth, into various fine chemicals holds significant value. Furfural stands out as one of the most promising platform compounds derived from lignocellulosic biomass. Due to its highly functional molecular structure, furfural can be selectively converted into various fuels and high-value compounds. This review discusses recent developments in furfural production and its conversion into related chemicals, such as furfuryl alcohol, γ-valerolactone, pentanediols, and nitrogen-containing compounds. It provides an in-depth understanding of the catalysts, systems, and mechanisms used in the selective transformation of furfural. The review also explores primary pathways and catalytic mechanisms, with a focus on advances in heterogeneous catalytic systems. Furthermore, it outlines future research directions and offers insights into potential applications in this field. This review presents several research trends, aiming to provide innovative ideas for further exploration of furfural downstream products in a greener, more efficient, and cost-effective manner.
Biomass is a unique form of energy wherein solar energy is captured and stored within green plants through photosynthesis, either directly or indirectly. It is Earth's exclusive and renewable carbon source, originating from a wide spectrum of materials and possessing substantial reserves.6 With its environmentally sustainable nature and rapid regeneration cycle, biomass can undergo chemical processes to produce a variety of valuable products.7 Consequently, biomass stands out as a pivotal choice in addressing the energy crisis.4,8–10 Leveraging renewable lignocellulosic biomass resources to generate high-value chemicals, like the conversion of various biomaterials such as straw and fruit shells into premium fuels and fine chemicals, emerges as a vital strategy in addressing future energy and environmental challenges. This approach has garnered growing attention from both academic and industrial sectors.11–14
In this context, furfural has garnered recent attention as a leading value-added chemical sourced from biomass. It is notably considered a key product in lignocellulosic biorefineries. Lignocellulosic biomass mainly refers to a class of natural products composed of cellulose, hemicellulose, and lignin, which can be hydrolyzed into C6 sugars (such as glucose and fructose) and C5 sugars (like xylose and arabinose) under the influence of acid catalysis.15 Furfural, recognized as one of the leading 30 platform compounds derived from biomass by the U.S. Department of Energy, is primarily obtained through the hydrolysis and dehydration of xylan, a prominent component of hemicellulose. The process of converting xylose/xylan into furfural is a thoroughly established method, with its initial industrial application dating back to 1921 when the Quaker Oats Company pioneered its commercialization. In general, raw materials containing a certain amount of pentose can be used to produce furfural. Due to the technical influence, high-valued utilization of furfural is still very limited. As a vital chemical raw material, furfural could serve various purposes, acting as a selective extractant, solvent, vulcanizing agent, anesthetic, intermediate for pesticides, a component in resin products, and a key ingredient for drug, cosmetic, and spice manufacturing. What's more, the inherent chemical reactivity of furfural, owing to its furan rings and aldehyde functional groups, positions it as a versatile biomass-derived platform compound. Realizing the high-value utilization of furfural and further expanding its applications in downstream, value-added products carry substantial scientific and industrial significance.16,17
Several excellent reviews on furfural conversion have been published, most of them focusing on biofuel applications.4,10,18,19 This article provides a systematic and comprehensive review of the progress of research on the catalytic transformation of furfural, a process leading to the production of valuable downstream products, such as furfuryl alcohol (FOL), tetrahydrofurfuryl alcohol (THFA), levulinic acid (LA), γ-valerolactone (GVL), 2-methylfuran (2-MF), 2-methyltetrahydrofuran (2-THMF), pentanediols (PeDs), cyclopentanone (CPO), and nitrogen-containing compounds with a special emphasis on recent advancements in heterogeneous catalytic material, green synthesis system and advanced technology in the last five years. Furthermore, possible reaction mechanisms for the conversion of furfural to a variety of high-value chemicals are discussed. Meanwhile, the challenges and opportunities for the upgrading of furfural to high-value chemicals are revealed, providing theoretical guidance for expanding the application value of furfural and pointing out the direction for further research in the future.
In industrial scale, furfural production usually consists of four main steps: hydrolysis of hemicellulose, dehydration of xylose, recovery of furfural, and finally sequential purification. At present, furfural industrial production usually uses batch reactors or continuous reactors. The whole furfural production process technology flow is about (Fig. 2):24 (1) first, corn cobs with no spoilage or mold were selected as raw materials for furfural production and crushed, as the natural moldy and deteriorated corn cobs would result in decreased polysaccharide content. After crushing, the particle size of the raw material is reduced, which increases the surface area of the raw material and thus the reaction speed. (2) Corn cobs with dilute acid are added and stirred well, then passed through a hydrolysis reaction in a hydrolysis pot. This process is difficult to operate at room temperature, so the actual production process is carried out at high temperature and pressure. The whole process temperature is controlled between 145 °C and 230 °C. Typically, high pressure steam is used to provide the heat required for the reaction, while another important role of steam is to remove furfural from the reaction system in time, thus avoiding the occurrence of side reactions. (3) Subsequently, distillation is performed to concentrate the dilute furfural solution, thereby enhancing the furfural concentration. After the dilute furfural solution is evaporated several times, the concentrated furfural vapors are led through the top of the tower, while the residual liquid is discharged from the bottom of the tower. The export steam enters the primary furfural collector through a condenser and divides into two layers, the upper layer is the saturated solution of furfural dissolved in water, containing 7–10% furfural content, and the lower layer of furfural is the concentration of raw furfural at about 90%. This process requires the temperature at the top of the tower to be maintained at 94–97 °C, the temperature at the bottom of the tower to be controlled at 98–103 °C, and the temperature of the distillate should be below 55 °C. (4) Due to the low purity of crude furfural, which contains high boiling point and low boiling point substances as well as water, its application is restricted and requires further refining. Usually, the refining is performed by decompression distillation or water vapor distillation, and the purity of furfural can be as high as 98% after the purification process.
Fig. 2 Schematic process flow diagram of typical furfural production from biomass.24 |
On the other hand, from the microscopic point of view of the chemical reaction, liquefaction of hemicellulose to xylose takes place in the following steps: (1) glycosidic bond protonation of xylan in acidic media. (2) The carbon–oxygen bond is cleaved to form carbon positive ions and monosaccharides, respectively. (3) The carbon positive ion binds to water and the positive charge migrates to eventually produce monosaccharides and protons. (4) The proton continues to participate in xylan liquefaction until the glycosidic bond reaction is complete.26
The reaction mechanism of xylose dehydration is very complex, and the following three different pathways are mainly considered according to different catalytic environments (Fig. 3): (i) the xylose is first ring-opened to form an enol under Lewis acid catalysis, then undergoes dehydration isomerization to form a diolefinic intermediate, and finally dehydration cyclization to form furfural. (ii) Catalyzed by Brønsted acid, the oxygen atom at the C1 or C2 position is protonated and undergoes dehydration and intracyclic rearrangement to form furfural directly. Or (iii) isomerization of xylose to form xylulose under the combined action of Lewis acid and Brønsted acid, which in turn undergoes dehydration cyclization to form furfural.27
Catalysts for furfural preparation can be divided into homogeneous catalysts and heterogeneous catalysts. The main types of homogeneous catalysts include common organic and inorganic acids, inorganic salts and metal ions, etc. The heterogeneous catalysts mainly refer to the solid acid catalysts, and the common solid acid catalysts mainly include molecular sieves, metal oxides, ion-exchange resins, phosphates and acidic heteropolyacids. The most widely used and technologically mature catalysts in the industry nowadays are H2SO428 and HCl,29 which catalyzed the production of furfural and can obtain high yield, but there are problems such as the difficulty of separating and recovering the products, the product contamination, and the loss of furfural yield due to the longer residence time. So, some people tried to use metal salt catalysts instead of acid catalysts. Leitner and María et al. have shown that a biphasic medium consisting of an aqueous solution of FeCl3 and NaCl with biomass-derived 2-THMF is an effective catalytic system for xylose dehydration.30 This method was able to achieve a maximum furfural yield of 71% in the presence of 20 wt% NaCl additive. Similar to the Fe system, AlCl3 catalyzed the conversion of xylose to >99% at 140 °C in the biphasic medium water/tetrahydrofuran (THF), with furfural yields up to 30% within 5 min.
Compared with homogeneous catalysts, heterogeneous catalysts can improve the conversion and selectivity of reactions by adjusting their acidity, acid amount and structure, thus becoming a hot research topic.10,31 Solid acid catalysts have many advantages in terms of activity, selectivity, catalyst life, and ease of recycling and reuse, and have been widely studied owing to they can directly replace liquid acids. O'Neill et al. used HZSM-5 catalyst to catalyze xylose in the aqueous phase at 200 °C to obtain 46% furfural yield.32 Yang et al. prepared a charcoal-based solid acid catalyst, SC-GCA-800, for the preparation of furfural by catalyzing the conversion of xylose in 1,4-dioxane using 4-anilinosulfonic acid as a sulfonating agent in the sulfonation of calcium gluconate, and achieved a furfural yield of 76.9% in 40 min at 140 °C.33 Qi et al. used cellulose as the char source, impregnated with FeCl3, and then sulfonated with sulfuric acid to produce a magnetic char-based solid acid catalyst, which effectively achieved the conversion of xylose to furfural in a water–methyl isobutyl ketone biphasic solvent system, reaching the highest yield of furfural of 79.0% at 190 °C.34 The development of heterogeneous solid acid catalysts has helped to increase furfural production, thus providing an opportunity for the furfural industry to explore more valuable downstream products.18,35
With the application of corn cobs as the raw material for production in China 30 years ago, production costs have been greatly reduced and China has gradually become the largest exporter of furfural and its processed products. In recent years, the furfural production process has become more and more perfect with the progress of social development and the growth of science and technology. Furfural has been vigorously promoted by many fields and industries, and its development prospects are very broad. However, there are still some problems in the actual production process, such as the environmental impact of the waste residue, hydrolysis pot wastewater and distillation tower top exhaust gas in the production process, the low efficiency of furfural production and so on. So the entire production process needs to be continuously developed and improved. It has improved the productivity while promoting growing sophistication of production process, manpower, material resources and other production costs to do an effective reduction in the overall development of agriculture to play a positive role in driving. At the same time, furfural and its derivatives require significant improvements in yield and production strategies to compete with petroleum-based products as renewable alternatives. The furfural industry is a basic industry for chemical products, and its production and demand are generally showing an upward trend from year to year. In China, for example, according to reported statistics, China's furfural production increased to 512200 tons in 2019 from 282000 tons in 2015, and demand increased to 493500 tons from 256500 tons in 2015, and it is expected that China's furfural production is expected to reach 600000 tons in 2023, and demand is expected to reach 590000 tons.36 Furfural production may be greater than demand, so it is necessary and important to study the conversion of furfural to other downstream products.
Furfural can be converted into a series of C4 and C5 molecules by a variety of catalytic processes, which are indispensable components in the production of liquid fuels and fuel additives (Fig. 4). Besides fuels and fuel additives, furfural can be converted into other valuable chemicals such as PeDs, GVL, butanediol and butyrolactone. Among them, FOL and THFA are simple hydrogenation products of furfural. Hydrogenolysis of the carbonyl group on the side chain of furfural after hydrogenation can produce 2-MF or 2-THMF.43 The furan ring in furfural can also be ring-opened and hydrogenated to produce PeDs. Reductive amination of furfural side-chain carbonyls to furfurylamine (FAM) and other C5 chemicals.4
In industrial production, hydrogenation is one of the important intermediate steps in the synthesis of high-demand products (flavors, pharmaceuticals, food preservatives, etc.). There are three ways to hydrogenate furfural: (1) Gas-phase hydrogenation: this is a traditional technology and the main reactor is a trickle bed reactor, which is a type of fixed-bed reactor. Hydrogen and feedstock are parallel flowed through a bed of solid particulate catalysts for gas–liquid–solid phase reaction. Trickle bed reactors commonly operate under pressure to promote mutual dissolution of the gas–liquid phases. Hydrogen and feedstock undergo full contact and hydrogenation reaction within the catalyst micropores. Furfural gas-phase hydrogenation is generally carried out at a temperature greater than 200 °C and under 3–6 MPa H2. In the gas-phase hydrogenation process, reaction activity and selectivity are not only affected by the catalyst, but also by the reaction temperature. Different temperature ranges produce different products, and higher temperatures make it easier to break the C–C bond.44 (2) Liquid-phase hydrogenation: this process involves dissolving sufficient hydrogen in the liquid-phase recycle material to provide for the hydrogenation reaction, so there is no need for a hydrogen recirculation system for the process. When hydrogen is dissolved in the feedstock oil and enters the reactor, mass transfer between the gas and liquid phases is avoided, the influence of the wetting factor is eliminated, the reactor temperature gradient is reduced, and the catalyst bed is close to isothermal operation. Moreover, the material flows unidirectionally through the catalyst bed, with a small temperature difference between the catalyst beds, fewer hydrocracking reactions, and higher liquid yields. At the same time, it makes the process flow simpler. The reaction conditions for liquid-phase hydrogenation are generally milder than those for gas-phase, with temperatures generally around 100–210 °C and H2 pressures generally 2–8 MPa.45 (3) Liquid-phase transfer hydrogenation: it uses a source of hydrogen obtained from a solvent or other additives. A more commonly used hydrogen source is isopropanol (2-PrOH), which is also as a solvent for the reaction. In the entire process, as the hydrogenation reduction reaction conflicts with the oxidation reaction, some of the reaction substrates may also be susceptible to oxidation. Therefore, liquid phase transfer hydrogenation usually requires nitrogen protection to exclude the influence of oxygen on the reduction reaction.46
In addition to the hydrogenation of the aldehyde group outside the furfural furan ring with the furan ring, there is also the possibility of hydrogenolysis reactions. Hydrogenolysis is usually a reaction in which a carbon–heteroatom bond (or carbon–carbon bond) is broken in a reduction reaction and hydrogen replaces the departing heteroatom (carbon atom) or group. Hydrogenolysis reactions are often accompanied by high temperatures and high pressures, under which the active metal is susceptible to sintering and coking, resulting in a decrease in catalytic activity.47 Metal oxide (e.g. ZrO2, TiO2, CeO2) loaded metal-based catalysts are a common form of catalyst combination. They can enhance the dispersion of metal particles while ensuring the catalyst's structural stability and resistance to loss of activity, and provide Brønsted acid or Lewis acid sites. This promotes the activation and breaking of carbon–heteroatom bonds for a more efficient hydrogenolysis process. In the hydrogenolysis reaction of furfural, in the presence of Brønsted acid or Lewis acid, breakage between C–O bonds occurs mainly. Breakage of the C–O bond outside the furan ring may yield 2-MF and 2-THMF. The ether bond on the furan ring may also undergo hydrolytic breakage to yield pentanediol or rearrangement after breakage in the presence of Brønsted acid to produce LA.
The aldehyde group in furfural can also be used in reductive amination reaction to produce amines, amines are used in a wide range of applications such as pharmaceuticals, rubber and pesticides.48 Among them, primary amine compounds are an important synthetic intermediates, which can be widely used in the production of plastics, resins and drugs, and so on.49,50 The typical methods for synthesis of primary amines include ammonolysis of haloalkane compounds, hydrogenation reduction of nitrile compounds, addition of olefins to ammonia, nitro hydrogenation reduction on benzene rings, and reductive amination of carbonyl compounds.51 Reductive amination of aldehydes and ketones for the synthesis of primary amines is one of the most widely used methods for the preparation of primary amines due to the advantages of easy operation, wide range of applicable reactions, and wide source of substrate aldehydes/ketones.52 Depending on whether or not intermediates such as imines need to be isolated during synthesis, carbonyl compound reductive amination reactions are classified into two types: indirect reductive amination and direct reductive amination. As the indirect reductive amination is usually free of competitive reduction reactions, the selectivity of the reductant is less required. But most of the imine intermediates are very unstable and difficult to isolate, direct reductive amination has been preferred.53 The synthesis of primary amines by direct reductive amination of aldehydes/ketones carbonyl compounds usually involves two steps: (1) imidization, the dehydration and condensation of aldehydes/ketones with ammonia donors (ammonia gas, liquid ammonia, and aqueous ammonia, etc.) to produce imine intermediates. (2) Sialylation hydrogenation, the hydrotreating of imines in the presence of a reducing agent to produce primary amines. The direct reductive amination of carbonyl compounds to synthesize high yield primary amines is challenging because many side reactions can be carried out simultaneously and by-products such as Schiff bases, secondary amines and alcohols are often present in the system.54
In recent years, researchers have focused on the study of new green and stable catalysts, including noble metal-based catalysts like Pt, Pd and Ru (Table 1),56,61,62 non-noble metals are mainly used Cu, Co and Ni (Table 2).63–68 The hydrogen sources used for hydrogenation are usually derived from molecular hydrogen (H2), methanol, 2-PrOH. The following is an overview of the catalytic reaction systems for the selective hydrogenation of furfural for the preparation of FOL and their latest research progress in recent years from the perspective of both noble and non-noble metal-based catalysts, respectively.
Entry | Catalysts | Solvent | T/°C | t/h | H-donor | FOL yield/% | Ref. |
---|---|---|---|---|---|---|---|
1 | Ptn/CoAl-MMOs | EtOH | 120 | 5 | H2 (0.8 MPa) | 98.7 | 69 |
2 | Pt1/CoAlOx | 2-PrOH | Room temperature | 24 | H2 (0.1 MPa) | 99.0 | 70 |
3 | Cu–Pt@TMS | 2-PrOH | 110 | 3 | H2 (1.0 MPa) | 99.5 | 71 |
4 | Pt/CeO2-270 | 2-PrOH | 80 | 1 | H2 (1.0 MPa) | 97.3 | 73 |
5 | Pt/HT | H2O | 28 | 2 | H2 (1.5 MPa) | >99.0 | 74 |
6 | Pd3PbMFs | 2-PrOH | 25 | 2 | H2 (0.1 MPa) | 93.0 | 75 |
7 | Ru3Co100 | H2O | 120 | 4 | H2 (0.1 MPa) | 100 | 78 |
8 | Pd–Cu/MCM-41 | 2-PrOH | 160 | 4 | 2-PrOH | 95.7 | 80 |
9 | Ru–Fe3O4/CNTs | 2-PrOH | 180 | 4 | 2-PrOH | 99.4 | 62 |
Entry | Catalysts | Solvent | T/°C | t/h | H-donor | FOL yield/% | Ref. |
---|---|---|---|---|---|---|---|
1 | CuNiOx(1/1) | 2-PrOH | 120 | 0.75 | H2 (3.0 MPa) | >97.0 | 81 |
2 | Cu3Co1/MgAlOx | 2-PrOH | 110 | 2 | H2 (2.0 MPa) | 99.9 | 82 |
3 | Ni@OMC | 1-PrOH | 180 | 4 | H2 (3.0 MPa) | 98.0 | 83 |
4 | POMs/Zr-MOF | 2-PrOH | 80 | 0.67 | 2-PrOH | 97.8 | 86 |
5 | Dx-MOF-808 | 2-PrOH | 90 | 2 | 2-PrOH | 94.4 | 87 |
6 | UIO-S0.6 | 2-PrOH | 160 | 3 | 2-PrOH | 94.7 | 88 |
Catalytic transfer hydrogenation has attracted great attention as a safe and environmentally friendly selective hydrogenation method for the preparation of high-value-added products from biomass resources in recent years. Use of hydrogen containing organics (especially cheap and abundant alcohol compounds) as hydrogen donors in the reaction, avoiding the use of H2 and thereby significantly reducing the complexity and cost of the experimental equipment and obtaining additional dehydrogenation products, has gradually attracted much interest in both industry and academia.79 Gao et al. prepared a series of bimetallic PdCu-loaded catalysts on MCM-41. Trace amounts of Pd (0.025 wt%) were introduced into the main body of Cu nanoparticles, resulting in a high distribution of Pd–Cu nanoparticles on the surface of the catalyst and in the pores of the MCM-41 supports. Compared with the high selectivity but poor activity of non-noble metal Cu and the high activity but low selectivity of precious metal Pd, the obtained Pd–Cu/MCM-41 catalysts showed excellent FOL selective hydrogenation performance. The reaction was carried out at 160 °C for 4 h with 2-PrOH as the hydrogen source, and 97.4% furfural conversion and 98.3% FOL selectivity were obtained.80 Li et al. used magnetic Fe3O4 modified Ru/carbon nanotubes (CNTs) catalysts (Ru–Fe3O4/CNTs) and reacted at 180 °C for 4 h using 2-PrOH as solvent and hydrogen donor to obtain FOL in 99.4% yield. After characterization, the catalyst active groups were analyzed to include the metal Ru0 and RuOx–Fe3O4 Lewis acid sites, which together synergistically contributed to the excellent performance of the catalyst.62
Cu-based, Ni-based, and Co-based catalysts can be used to catalyze the synthesis of FOL from furfural hydrogenation. Fang et al. synthesized a mesoporous CuNiOx catalyst using KIT-6 as a template, and the reaction was carried out at 120 °C for 45 min under 3 MPa H2 conditions to obtain a FOL yield of >97%.81 The combination of experimental characterization and density-functional theory calculations reveals the key role of Cu+ matter in hydrogen activation. In addition, the CuNi alloy phase is considered to be an adsorption site for furfural, and the interaction between the Cu+ site and the CuNi alloy phase enhances the catalytic activity of the CuNi catalysts in furfural hydrogenation as compared to the monometallic Cu and Ni catalysts. Zhao et al. prepared a series of layered double hydroxide (LDH)-derived CunCo1/MgAlOx (n = 1, 2, 3, and 4) catalysts by a one-pot method. Among them, the synergistic effect of Cu and Co significantly increased the catalyst activity and favored the formation of Cu2O, which could act as a Lewis acid site to polarize the carbonyl group, while in situ DRIFTS showed that Cu3Co1/MgAlOx preferentially adsorbed and activated CO rather than the furan ring, thus reducing the occurrence of side reactions to promote the formation of FOL. The reaction was carried out at 110 °C and 2 MPa H2 for 2 h, achieving 99.9% of FOL yield.82 Tang et al. prepared Ni-doped ordered mesoporous carbon (Ni@OMC) by a one-pot solvent evaporation induced self-assembly process. Due to the coordination between gallic acid and Ni2+, the Ni particles were confined in the OMC carbon skeleton and could be uniformly dispersed in the carbon skeleton. The synthesized Ni@OMC was reacted at 180 °C and 3 MPa H2 for 4 h, providing a yield of FOL as high as 98%.83
As an emerging porous material, metal–organic frameworks (MOFs) have attracted much attention due to their unique advantages such as controllable composition, large specific surface area, high porosity.84,85 Based on the tunability of MOF precursors, it is possible to modify the organic or inorganic part of the structure from the molecular level, or to load multiple target substances into the pore structure of the MOF precursor and combine them with other materials, resulting in the formation of a variety of multi-component and multi-functional composite derivatives, which possess certain physical properties of MOF materials and can form synergistic effects within them. Combining multiple functional groups for synergistic effects is a method for designing multifunctional catalysts. Han et al. reported a bimetallic cluster catalyst (POMs/Zr-MOFs) that grafts polyoxometalate single clusters onto Zr–O clusters of Zr-MOFs via synergistic acid–base couple site and modulates the metal nodes of the Zr–O clusters for enhanced active site accessibility. The porosity and dual active centers of the POMs/Zr-MOFs make the substrate more accessible to the catalytic site, providing more favorable conditions for the catalytic transfer hydrogenation of furfural.86 The reaction was carried out at 80 °C for 40 min with 2-PrOH as the hydrogen donor, and the FOL selectivity reached 97.8%. Fu et al. synthesized defective MOF-808 (Dx-MOF-808) and obtained catalysts with defects highly exposing Zr sites by introducing benzoic acid as a temporary ligand to compete with the formal ligand 1,3,5-benzenetricarboxylic acid for coordination with the metal clusters and then removing the temporary ligand with methanol.87 The results showed that benzoic acid could not only participate in the synthesis of MOF-808 as a temporary ligand to obtain defective structures, but also act as a moderator to control the particle size of the catalyst. The reaction was carried out at 90 °C for 2 h with 2-PrOH as the hydrogen donor, and the FOL selectivity reached 94.4%. Wu et al. prepared sulfonic acid-functionalized microporous zirconium metal–organic skeletons (UIO-Sx, x represents the molar fraction of sulfonic acid-containing), and the modification of sulfonic acid groups on UIO-66 resulted in the formation of stronger Lewis acid/base and Brønsted acid sites.88 The reaction was carried out in 2-PrOH at 160 °C for 3 h, exhibiting 94.7% FOL yield. Kinetic experiments showed that the activation energy of catalytic transfer hydrogenation of furfural was as low as 50.8 kJ mol−1 over the UIO-S0.6 catalyst.
In transfer hydrogenation, the hydrogenation performance of the catalysts decreased slightly with the increase in the length of the carbon chain of the secondary alcohols. This may be due to the negative steric effect caused by the longer side chains of secondary alcohol hindering the formation of transition states between furfural and the active site. In general, if tert-butanol is the solvent, the catalyst reflects less catalytic transfer hydrogenation effect than that in solvents such as ethanol, 2-PrOH, 2-butanol, etc., can be used as a hydrogen source, or even the reaction cannot begin. This suggests that the presence of α-H or β-H in alcohols is crucial for the MPV reduction reaction.
Entry | Catalysts | Solvent | Reaction conditions | THFA yield/% | Ref. |
---|---|---|---|---|---|
1 | Pd1Cu9 | H2O | 2 mmol furfural, 0.04 mmol catalyst, 20 mL surfactant/H2O, 1.0 MPa H2, 50 °C, 4 h | 95.6 | 77 |
2 | PdNiCo/N-CNTs | EtOH | 10 mL EtOH furfural solution, 0.05 g catalyst, 120 °C, 3.0 MPa H2, 3 h | 97.1 | 91 |
3 | Pt(3)Ni(3)/C | H2O | 0.2 g furfural, 0.1 g catalyst, 20 mL H2O, 35 °C, 2 MPa H2, 12 h | 93.0 | 92 |
4 | Ni/MMO–CO3 | 2-PrOH | 0.5 mL furfural, 0.1 g catalyst, 30 mL 2-PrOH, 110 °C, 3 MPa H2, 3 h | 99.0 | 94 |
5 | Co/CoOx@N-CNTs | EtOH | 200 mg furfural, 50 mg catalyst, 30 mL EtOH, 150 °C, 3.0 MPa H2, 10 h | 99.5 | 95 |
6 | Ni/C-400 | EtOH | 0.6 g furfural, 0.1 g catalyst, 30 mL EtOH, 80 °C, 1 MPa H2, 4 h | 98.5 | 96 |
7 | Ni-CA-400 | 2-PrOH/H2O | 0.48 g furfural, 25 mg catalyst, 9 mL 2-PrOH, 1 mL H2O, 80 °C, 3.0 MPa H2, 3 h | 99.8 | 97 |
Noble metal catalysts tend to be highly active, not easily lost, and require mild reaction conditions. For the simultaneous hydrogenation of CC and CO double bonds, the use of multi-metal co-catalysis is also an important idea. Synthesis of bimetallic or polymetallic central catalysts by doping non-noble metal catalysts into noble metal catalysts is the current research hotspot for furfural total hydrogenation catalyst design. The Pd-based bimetallic alloy catalyst (Pd1Cu9), prepared by the previously mentioned Chen team by reducing a metal precursor dissolved in an aqueous solution of the sugar-based surfactant GluM, efficiently converts furfural to FOL.77 Replacing Cu with Ni in the bimetallic, the Pd1Ni3 alloy prepared can catalyze the full hydrogenation of furfural to generate THFA with high selectivity, and the reaction was carried out for 4 h at 50 °C and 1 MPa H2 with GluM aqueous solution as solvent, and the selectivity of THFA was 96.2%. Wu et al. prepared Pt(3)Ni(3)/C catalysts by distributing small particles of PtNi alloy on activated carbon. Electron-rich metal singlet Pt promotes hydrogen dissociation and facilitates the hydrogenation of furfural, and the furan ring adsorbs on the Ni surface, which together synergistically enhance the catalytic effect.92 The furfural conversion was 99% and the THFA yield was 93% after 12 h of reaction at 35 °C and 2 MPa H2 with H2O as solvent. Ruan et al. prepared PdNiCo/N-CNTs catalysts by loading trimetallic PdNiCo on nitrogen-doped carbon nanotubes, which showed good catalytic performance due to its own PdNiCo alloy nanostructures as well as electronic synergistic effects between PdNiCo metal nanoparticles and N-CNTs. THFA could be obtained in 97% yield by reacting in ethanol at 120 °C and 3 MPa H2 for 3 h.91
Recently, researchers have also focused on the design of catalysts involving only non-noble metals in an attempt to improve the yield by changing the catalyst structure and adjusting the adsorption configuration of furfural on the catalyst. In general, nickel species are active on both the furan ring (CC) and the carbonyl group (CO) of the furfural molecule, whereas copper species are highly active only on the carbonyl group of the furfural molecule,93 and both non-noble metals are also commonly used in furfural hydrogenation reactions. These two types of non-noble metals are also commonly used in furfural hydrogenation reactions. At present, non-noble metal catalysts have not been applied to the direct transfer hydrogenation of furfural to THFA, but there have been many studies on the use of non-noble metal catalysts for gas-phase hydrogenation. Meng et al. utilized metal nitrates or carbonates to prepare two catalysts with highly exposed Ni(111) facets loaded on polymetallic oxides with hydrotalcite as the precursor. The Ni/MMO–CO3 showed the only selectivity for THFA, with 99% yield in 2-PrOH at a reaction temperature of 120 °C and 3 MPa H2 for 3 h. The reaction temperature of Ni/MMO–CO3 was also shown to be the only selective agent for THFA. The characterization revealed that terrace sites are predominant on Ni/MMO–CO3 surface, which gives furfural a parallel adsorption conformation on the catalyst and facilitates the activated adsorption of CO groups and CC bonds in the furan ring.94 Ranaware et al. used cobalt and cobalt oxides loaded on N-doped carbon nanotubes over catalysts (Co/CoOx@N-CNTs), the absence of Brønsted acid sites in ethanol and the involvement of three hydrogen atoms in the ethanol molecule led to the predominance of THFA as a major product during furfural hydrogenation, and under optimal reaction conditions, the yield of THFA could achieve 99.5%.95 Fu et al. employed 3D floral MOF Ni-BDC as a sacrificial template to prepare a carbon-embedded nickel-based catalyst Ni/C-400. With a highly exposed Ni(200) facet and highly dispersed surface NiO species, the catalyst exhibited excellent catalytic performance for the total hydrogenation of furfural, achieving an excellent THFA yield of 98.5% at 80 °C and 1 MPa H2. The NiO species on the surface has a greater ability to adsorb furfural than Ni particles, while the Ni(200) surface is able to adsorb more dissociated H atoms, which allows the catalyst to achieve both the key to high conversion and high selectivity.96 Sheng et al. synthesized precursors of nickel citrate complexes using citric acid as a carbon source and complexing agent, and subsequently prepared Ni-based catalysts (Ni-CA-400) with carbon-covered structure by pyrolysis, which possessed a large specific surface area, abundant defective sites in the carbon layer, and uniform Ni nanoparticles. In furfural hydrogenation reaction, Ni-CA-400 gave a high yield of 99.8% THFA at 80 °C under mild conditions. Such excellent catalytic performance is attributed to the unique carbon-coated structure of the catalyst, which not only effectively avoids the migration and aggregation of the active nickel particles during the preparation process, but also protects the metallic nickel nanoparticles from oxidative exposure to the air, which can lead to catalyst deactivation.97
From the above studies, it can be seen that the structure of the catalyst and the size of the catalyst active center have an important influence on the type of adsorption of the reaction substrate, which leads to different selectivity of the hydrogenation reaction. Therefore, identifying the catalytic properties of each active site is crucial for designing hydrogenation catalysts. Meanwhile, alloy structures with high catalytic surface area and tunable surface energy can lead to high catalytic selectivity and activity. How to control the composition of bimetallic alloys to regulate furfural hydrogenation to generate high-value-added derivatives is an important strategy for catalyst design, and a great challenge in catalyst design at present.
Xu et al. investigated the mechanism of furfural hydrogenation on Ptn/CoAlOx (Fig. 5a). Molecular hydrogen first adsorbed on the metal sites after activation to activated hydrogen and then added to the subsequent reaction. Due to the differences in the gathering of Pt species, furfural has different adsorption modes on Pt single atoms or Pt clusters. The formation of the final product of the furfural hydrogenation reaction is largely influenced by the adsorption selectivity of the CC and CO bonds. The CO of furfural can be adsorbed on both single-atom and clusters, but the CC bond in the furan ring can only be adsorbed on Pt clusters. The selective adsorption of the CO bond on single-atom leads to the unsaturated FOL product, whereas, because of the simultaneous adsorption of CC and CO, the saturated hydrogenation product, THFA, is generated on Pt nanoclusters.70 Ranaware et al. investigated the mechanism of THFA generation from furfural hydrogenation on Co/CoOx@N-CNTs (Fig. 5b and c). The O of the aldehyde group in the furfural molecule is adsorbed on the Coδ+ center. Furfural was selectively hydrogenated to FOL by H ions dissociated from H2 at the metal Co0 site and H+ ions dissociated from ethanol on the OV–Co3O4 surface. Subsequently, FOL was adsorbed on the CoOx surface and H ions dissociated from H2 and ethanol saturated the furfuryl ring to produce THFA.95
Fig. 5 Plausible reaction mechanisms of (a) furfural hydrogenation to FOL or THFA over Pt1/CoAlOx or Ptn/CoAlOx, (b and c) furfural hydrogenation to THFA over Co/CoOx@N-CNTs in ethanol, (d) catalytic transfer hydrogenation of furfural.62,70,95 |
In summary, the mechanism of furfural hydrogenation for the preparation of FOL can be summarized as follows:62 (1) in metal-mediated hydrogenation, activated H atoms are first adsorbed onto the metal sites, and then the adsorbed H atoms are added to the CO bond to form an O–H bond; (2) whereas in the Lewis-acid-mediated transfer hydrogenation with alcohols as the hydrogen donor as well as the solvent is transferred according to the MPV mechanism. The MPV mechanism starts with the Lewis acid sites. The adsorbed 2-PrOH and furfural on the L acid site form a propanol–furfural six-membered ring configuration, through which the β-H of the 2-PrOH is transferred to the carbonyl C atom of the furfural (Fig. 5d).
In addition to the commonly assumed pathway of THFA via the carbonyl group of furfural firstly hydrogenated to produce FOL followed by further hydrogenation of the furan ring, Huang et al. found that there is another pathway from furfural hydrogenation to THFA via the THFAL intermediate followed by carbonyl group hydrogenation to produce THFA (Fig. 6).98 Huang estimated the activation energy for each step experimentally and computationally. Since Ru/ZrO2 is well known as an active catalyst for CO reduction, the apparent Ea of furfural over this catalyst after the first step of pathway 1 was calculated to be about 56 kJ mol−1. The second step of pathway 1 was then tested on a Pd/Al2O3 catalyst and the apparent Ea was found to be about 44 kJ mol−1. This suggests that the decisive speed step in pathway 1 is the reduction of the CO bond. Using the same method, the apparent Ea of the first step of pathway 2 was first tested on Pd/Al2O3, and the apparent Ea of furfural to THFAL was about 21 kJ mol−1. Then after a kinetic study, the apparent Ea for the hydrogenation of THFAL to THFA was tested and obtained to be 30 kJ mol−1. From this, Huang deduced that the linkage structure of the furan ring and the carbonyl group may reduce the hydrogenation activity of the CO bond, and that the hydrogenation of the CO bond of THFAL overcomes a much lower energy barrier than that of pathway 1. Of course, for rapid reactions with such low activation energies, the observed conversion may be affected by mass transfer limitations rather than catalyst activity, which would mask the actual differences between different catalysts, leading to false conclusions. This is also something that researchers need to pay more attention to in the studies.
Fig. 6 Pathways for hydrogenation of furfural to THFA.98 |
The proposal of two pathways and the kinetic studies provided a new understanding of the multimetallic synergistic catalytic hydrogenation.
Entry | Catalysts | H-donor | Reaction conditions | LA yield/% | Ref. |
---|---|---|---|---|---|
1 | 1.5 wt% Cu/NbP | H2O/HCOOH | 1 mmol furfural, 50 mg catalyst, 5 mL H2O, 10 mL HCOOH, 160 °C, 3 h | 67.0 | 102 |
2 | HTC–SO3H | GVL/H2O | 3 g furfural residue, 1.5 g catalyst, 45 mL GVL/H2O (1:1, v/v), 190 °C, 3 h | 74.8 | 103 |
3 | Amberlyst-15 | H2O | Catalyst:formalin:furfural = 2:2:1 w/w/w, 2 g H2O, 160 °C, 5 min | 44.4 | 104 |
Zhang et al. used carbonization–grafting–oxidation method to prepare a bifunctional carbon carrier solid acid catalyst (HTC–SO3H) containing both cellulose binding sites (–Cl, –COOH and –OH) and catalytic sites (–SO3H). These groups (–Cl, –COOH, and –OH) can form molecular hydrogen bonds with cellulose or oligosaccharides, promote the binding of solid acids to feedstocks, and can effectively hydrolyze furfural residues to LA. In addition, Fe3O4 was added to synthesize a magnetic bifunctional solid acid catalyst (HTC–SO3H–Fe3O4), and its catalytic performance was investigated. After 3 h of hydrolysis in γ-pentolactone–H2O (1:1, v/v) at 190 °C, the total yield of LA was 74.81% (HTC–SO3H) and 70.18% (HTC–SO3H–Fe3O4), respectively. The water in the solvent promoted the production of LA and GVL facilitated the dissolution of cellulose. This study provides a green and feasible way to prepare biomass hydrolysis catalysts with high catalytic activity and high recovery.103
Velaga et al. reported a novel green reaction route for the production of LA via furfural in a microwave (MW) reactor. Using a pure Amberlyst-15 catalyst in a MW reactor with a formaldehyde:furfural mass ratio of 2:1 at 160 °C. Surprisingly, the reaction was almost complete within 30 min, and after a shorter time (5 min) furfural showed an increase in side reactions with formaldehyde. Under the reaction condition, the conversion of furfural was 74% and the LA selectivity reached 60%. The LA yield decreased by a factor of 2.7 as the furfural concentration increased from 10 wt% to 50 wt% at constant catalyst loading. After kinetic/thermal effect analysis, MW heating increased the FF conversion and LA yield by a factor of 3.5 and 4.6 compared to conventional heating. The process differs from the normal furfural pathway for LA production in that formaldehyde is added to the reaction system, and furfural first reacts with formaldehyde via hydroxymethylation to form the reaction intermediate 5-hydroxymethylfurfural (HMF), and then undergoes a ring-opening process of dehydration and decarbonylation to produce the final product LA. The generation of the reaction intermediate HMF avoids the need to use high-pressure hydrogen and expensive metal catalysts in the hydrogenation step.104
In catalyst for the preparation of LA from furfural multiphase catalytic one-pot method, the regulation of the ratio of the acid amount of Lewis acid and Brønsted acid as well as the suitability is the key to the success of efficient catalyst preparation, which will directly affect the catalytic effect of the reaction, the reuse and recycling effect of the catalyst. Impregnation loading of Lewis acid and Brønsted acid, ion exchange doping of active centers, etc. are major methods to modulate the amount of Brønsted acid and Lewis acid in catalysts currently. Future research should also vigorously explore other methods to efficiently regulate and construct catalysts with appropriate amounts of highly active Brønsted acid and Lewis acid active centers.
The breaking of the C–O bond in furan ring is the key point in the conversion of furfural to LA. The hydroxyl group of FOL is protonated in the presence of Brønsted acid and forms a six-membered ring transition state assisted by a solvent molecule, after which the rearrangement completes the breaking of the C–O bond in the furan ring. The C–O bond in the furan ring is broken in the presence of the solvent molecule. The free energy of activation barrier in this process is higher (21.3 kcal mol−1).105 Therefore, increasing the temperature is thermodynamically more favorable for the breakage of the C–O bond in the furan ring, while the applied pressure added outside the system compartment also favors the breakage of the C–O bond in the furan ring. In addition, various factors such as the surface morphology, pore structure, pore size and the choice of hydrogen donor of the catalysts will also affect the catalytic performance of the bifunctional catalysts in catalyzing the conversion of furfural to LA, which are all issues that need to be taken into consideration in designing the highly efficient bifunctional catalysts for the one-step conversion of furfural to LA.
The Meerwein–Ponndorf–Verley (MPV) reduction reaction pathway allows the use of inexpensive metal oxides with Lewis acid/base properties instead of precious metal catalysts, and the use of secondary alcohols as hydrogen donors instead of high-pressure H2 enables selective hydrogenation of LA or furfural to the corresponding products via the transfer hydrogenation pathway.110–112 Solid acids, such as H-type zeolites (HZSM-5), heteropoly acid and even metal oxides (e.g., γ-Al2O3) are very active in the ring-opening reaction of furan rings. Hydrothermally stable molecular sieves such as beta zeolite and ZSM-5 zeolite with good mass transfer effect are commonly used as carriers for this reaction, and the introduction of active centers such as Zr, Hf compounds with Lewis acid sites and phosphotungstic acid (HPW) with Brønsted acid sites are commonly used for designing highly efficient bifunctional catalysts (Table 5).113–120
Entry | Catalysts | H-donor | L acid/B acid | Reaction conditions | GVL yield/% | Ref. |
---|---|---|---|---|---|---|
1 | HPW/Zr-beta | 2-PrOH | 8.0 | 0.11 mmol furfural, catalyst/furfural = 0.4, 20 mL 2-PrOH, 160 °C, 24 h | 68 | 113 |
2 | FM-Zr-ARS | 2-PrOH | 2.1 | 1 mmol furfural, 0.1 g catalyst, 5 mL 2-PrOH, 160 °C, 8 h | 72.4 | 115 |
3 | 10% TPA/20% ZrO2-SBA-15 | 2-PrOH | 0.77 | 0.196 g furfural, 7.5 g cat. per L, 20 mL 2-PrOH, 170 °C, 12 h | 81 | 122 |
4 | Zr-CN/H-β | 2-PrOH | 0.45 | 0.22 g furfural, 0.1 g catalyst, 5 mL 2-PrOH, 160 °C, 18 h | 76.5 | 123 |
5 | Zr-P/SAPO-34 | 2-PrOH | 2.0 | 1 mmol furfural, 0.2 g catalyst, 20 mL 2-PrOH, 150 °C, 18 h | 80.0 | 126 |
6 | Fe3O4/ZrO2@MCM-41 | 2-PrOH | 8.3 | 0.5 mmol furfural, 40 mg catalyst, 10 mL 2-PrOH, 150 °C, 30 h | 85 | 127 |
7 | Sulfate DUT-67 (Hf)-0.06 | 2-PrOH | 2.5 | 1.25 mmol furfural, 0.3 g catalyst, 25 mL 2-PrOH, 180 °C, 24 h | 87.1 | 129 |
8 | VPA-Hf(1:1.5)-0.5 | 2-PrOH | 2.1 | 1 mmol furfural, 0.2 g catalyst, 10 mL 2-PrOH, 180 °C, 14 h | 81 | 131 |
Zhang et al. achieved one-pot conversion of furfural to GVL using a physical mixture of solid Zr-HY Lewis acid and Al-HY Brønsted acid catalysts, in which the ratio and strength of the Lewis and Brønsted acid sites could be independently adjusted. Zr-HY zeolites have larger pore sizes and stronger Lewis acid sites compared to Zr-beta. As Brønsted acid sites, Al-HY zeolites are more effective than Al-beta because they convert furfuryl ethers to the intermediate levulinate esters (LEs) instead of β-angelicalactone. The mixed catalysts of Zr-HY (Si/Zr 20) and Al-HY (Si/Al 6) showed very significant catalytic activity and 85% GVL yield after 5 h of reaction at 120 °C.121 Srinivasa Rao et al. prepared a series of metal oxide and HPW catalysts loaded on β-zeolite catalysts by impregnation method, with β-zeolite containing 20% ZrO2 and 5% HPW showing high activity for the reaction, and the GVL yield was 85% in 2-PrOH at 170 °C for 10 h. The high Lewis acid density as well as the basic sites are responsible for the good catalytic activity of the catalyst.122 Zhang et al. prepared Zr-graphite carbon nitride/H-β complex (Zr-CN/H-β) as a catalyst for the reaction by impregnation–pyrolysis method. When 2-PrOH was used as the solvent and the reaction was carried out at 160 °C for 18 h, the GVL yield was 76.5%. The introduction of both Zr species and carbon nitride into H-β in the catalyst formed novel acids and bases that should be responsible for the efficient conversion of furfural to GVL.123 Kim et al. synthesized ZrO2-containing mesoporous Al-MFI nanosponges (ZrO2[Al]MFI), the unique mesoporous structure of which aggregated zirconia clusters (Lewis acid sites) onto the outer surface of the crystals, whereas leading to the separation of Brønsted acid and Lewis acid sites on the inside and outside of the support, respectively, and thus exhibited high catalytic activity. The GVL yield could reach 82.8% after 36 h of reaction at 170 °C.124 Similarly, Rao et al. prepared catalysts to separate Brønsted acid sites from Lewis acid sites and showed that the catalysts with ZrO2 present in the SBA-15 pores and HPW dispersed on the support exhibited the highest activity, with complete conversion of furfural and 81% GVL yield for 11 h of reaction at 170 °C.125 Li et al. prepared a series of zirconium phosphate (ZPS-X) catalysts loaded by SAPO-34. The introduction of zirconium phosphate resulted in the formation of strong Lewis acid sites in the catalysts, and the catalytic activity of the ZPS-X catalysts for the reaction could be adjusted by adjusting the molar ratio of Zr to P (Zr/P ratio) and changing the number of Lewis and Brønsted acid sites. When the Zr/P ratio was adjusted to 1.0, the furfural was completely converted by the reaction at 150 °C for 18 h, and the yield of GVL reached 80.0%.126 Gao et al. prepared a multi-functional catalyst Fe3O4/ZrO2@MCM-41 by impregnation, and the doping of Fe3O4 not only endows the catalyst with strong magnetism, which is favorable for the recycling of the catalyst, but also regulates its acidity to promote the production of GVL. After 30 h of reaction at 150 °C, the GVL yield reached 85%. As well, a kinetic study illustrated that the alcoholysis of the LEs is the rate-limiting step of the whole process.127 Liu et al. prepared a series of bifunctional catalysts Zr-SBA-15-x with different Si/Zr molar ratios and independent Lewis and Brønsted acid sites by in situ synthesis. When the Zr loading was increased appropriately, the GVL selectivity increased and a lower selectivity of isopropyl 4-oxopentanoate (IPL) was obtained. The reaction in 2-PrOH at 190 °C for 24 h resulted in a GVL selectivity of 93.3%.128
Li et al. synthesized a novel bifunctional catalyst sulfate DUT-67 (Hf) by modifying the Hf-based metal–organic skeleton. The acidity of the catalyst was adjusted by immersion in aqueous sulfuric acid solutions of different concentrations to modify the molecules of zirconium clusters. DUT-67 (Hf) acidified by 0.024 mol L−1 aqueous sulfuric acid solution exhibited the best catalytic activity, with a yield of 84.9% of GVL after 20 h of reaction at 180 °C under the synergistic effect of Lewis and Brønsted acid.129 Tang et al. have developed a hierarchically structured bifunctional Hf-Al-USY zeolite with balanced Brønsted and Lewis acid sites, and through the synergistic effect of hydrolytic ring-opening reaction of Brønsted acid at the Al site and H-transfer hydrogenation of Lewis acid via the Hf site as well as the improved mass transfer via the de-alumination-induced hierarchical structure, the catalyst showed a very high performance in the reaction, with the GVL yield at 140 °C for 20 h. The GVL yield was 81%.119 Antunes et al. prepared Hf-containing beta-zeolite materials with intracrystalline hierarchical pore structures from a top-down approach for the first time for the production of furfural one-pot conversion to GVL. The catalytic effect and development potential of the catalysts were evaluated by kinetic modeling studies, solid-state spectroscopic characterization, and catalyst stability studies. The most effective catalyst was the hierarchical (intracrystalline micro/mesopores) microcrystalline material Hf-WdeSAlBeta-m, which means Hf-containing desiliconized and dealuminized microcrystalline zeolites. After 24 h of reaction at 180 °C, the GVL yield was 73%, which was related to its stronger acidity and mesopore size and porosity. This is the first report of Hf-containing beta-zeolite with intracrystalline layered pore structures, and the results highlight the potential for the development of these types of materials in GVL production.130 Tan et al. prepared a variety of novel coordination organic phosphate-Hf polymers functionalized with Brønsted/Lewis acids and alkali sites using a one-pot solvothermal method from vinylphosphonic acid (VPA), p-toluenesulfonic acid (p-TSA), and HfCl4. The intensity and content of Lewis and Brønsted acid could be adjusted by adjusting the molar ratio of VPA, p-TSA and HfCl4. The prepared VPA-Hf(1:1.5)-0.5 catalyst has a spherical porous structure with large surface area and pore volume, and can achieve 81.0% GVL yield under optimized conditions.131
Based on the above studies, the difference in catalyst reactivity is affected by the number of active centers and is related to the relative content of Brønsted and Lewis acid. However, an excess of Brønsted acid in the system will cause the reaction of furfural, furfuryl ether and β-angelicalactone to form the inert compound humin, which may cover the catalysts’ active site and reduce or deactivate the catalyst activity.132 Thus, it can be seen that the balance between Brønsted and Lewis acid sites is the key to improving the catalytic performance. Meanwhile, the catalytic performance of the catalyst is also closely related to the support structure. In general, a large specific surface area and porosity will provide the reactants with more substrate-accessible active sites, and a large channel or cage structure helps the diffusion of reacting molecules, making it easier for the reactants to reach the active sites.
Fig. 8 Possible reaction mechanism for the conversion of furfural to GVL.133 |
The whole process is a series of cascade reactions, such as catalytic transfer hydrogenation reaction, etherification reaction, ring-opening reaction, partial hydrogenation reaction and cyclization reaction, which is very complex, and this complex reaction process requires higher performance catalysts. According to the existing studies, the difference in catalyst reactivity is influenced by the number of active centers, which is related to the relative content of Brønsted and Lewis acids. The excess of Brønsted acid in the system, however, causes the reaction of furfural, furfuryl ether and β-angelicalactone to form the inert compound humin, which may cover the catalyst active site and reduce or deactivate the catalyst activity.132 Thus, it can be seen that the balance between Brønsted and Lewis acid sites is the key to improving the catalytic performance. Meanwhile, the catalytic performance of the catalyst is also closely related to the carrier structure. In terms of catalysis, the large specific surface area and porosity will provide more substrate accessible active sites for the reactants, and the large channel or cage structure helps the diffusion of the reacting molecules, making it easier for the reactants to reach the active sites. Therefore, it is necessary to consider the structure and acid/base properties of the catalyst when designing the catalyst in order to improve the activity of the catalyst.
Although the development of efficient catalytic systems for the conversion of furfural to GVL has been greatly advanced, a number of challenges still exist. For example, in the one-pot conversion of furfural to GVL, due to the high activation energy required for the FOL ring-opening reaction, it needs to be carried out successfully at higher temperatures, but humins is easily produced under these conditions, reducing the yield. Therefore, how to overcome the high temperature and yield problems and design new and efficient catalysts is still the current difficulty.
Entry | Catalysts | Solvent | T/°C | t/h | H-donor | 2-MF yield/% | Ref. |
---|---|---|---|---|---|---|---|
1 | NiMo IMC | 2-PrOH | 200 | 6 | H2 (0.1 MPa) | 99.0 | 155 |
2 | Ni/Fe3O4 | 2-PrOH | 100 | 4 | H2 (1 MPa) | 78.4 | 156 |
3 | NiCo–MgAlO | 2-BuOH | 220 | 6 | H2 (3 MPa) | 92.3 | 140 |
4 | Cu/CuFe2O4@C-A | 2-PrOH | 165 | 5 | H2 (1.5 MPa) | 90.0 | 159 |
5 | Cu–Mo/CoOx | 2-PrOH | 180 | 4 | H2 (2 MPa) | 92.0 | 144 |
6 | CuCo/NC | 1,4-Dioxane | 150 | 4 | H2 (1.5 MPa) | 95.7 | 158 |
7 | Co/CoAl2O4 | EtOH | 150 | 5 | H2 (1.5 MPa) | 97.0 | 161 |
8 | CuZnAl | 2-PrOH | 180 | 4 | 2-PrOH | 72.0 | 160 |
9 | CuFe2O4 | 2-PrOH | 200 | 1.5 | 2-PrOH | 97.0 | 162 |
Zhou et al. prepared a 5Cu3Re/Al2O3 bifunctional catalyst using a conventional impregnation method. Using isopropanol as the hydrogen donor, 94.0% yield of 2-MF was achieved at 220 °C for 4 h. The excellent performance of the bimetallic CuRe/Al2O3 catalysts can be attributed to the synergistic effect of the Al2O3 carrier, metal Cu and Re species. In particular, Al2O3 has abundant acid–base sites and large specific surface area, and the dehydrogenation of isopropanol and the hydrogenation of furfural to FOL can be enhanced with the introduction of Cu on Al2O3 surface. The modification of Re species, on the other hand, promotes the further hydrogenolysis of FOL to 2-MF. In addition, the strong interaction of ReOx with Cu and Al2O3 results in the high stability of the bimetallic catalysts. The selectivity of 2-MF and the catalyst activity did not change obviously during five cycles.151
The above studies have brought some inspirations for future development and research on noble metal catalysts. (1) Developing more catalysts for green solvents such as H2O, isopropanol, etc. (2) Doping non-precious metals to construct bifunctional catalysts to reduce the cost and at the same time to realize the improvement of catalytic performance. (3) Designing the catalyst structure on the nanoscale and constructing the active sites on the atomic level.
Liu et al. prepared a NiMo intermetallic compound catalyst (NiMo IMC) based on a layered bimetallic hydroxide precursor, which exhibited outstanding catalytic performance for the hydrodeoxygenation of furfural to 2-MF. 99% yield of 2-MF was achieved by reacting at 200 °C for 4 h at a rather low hydrogen pressure (0.1 MPa).155 Characterization showed that the atomically-ordered Ni and Mo sites in the catalyst determined the furfural molecule exhibits a unique adsorption pattern, with the CO group binding to adjacent Ni–Mo sites (via Mo–O and Ni–C bonds). Furan rings exhibit weak, tilted physisorption on the NiMo IMC(331) surface through the Ni–C bond. This unique configuration promotes the activated adsorption of CO groups, whereas the adsorption of the furan ring is greatly inhibited, which leads to the inhibition of saturated hydrogenation of the furan ring and the breakage of the C–O bond outside the ring. Wang et al. also synthesized the catalyst Ni/Fe3O4 from layered bimetallic hydroxides. When the Ni/Fe ratio was 1, the yield of 2-MF reached 78.4% under mild conditions (100 °C, 1 MPa H2, 4 h).156 Characterization revealed that the Ni2+/Fe3+ ratio was closely related to the dispersion and surface acidity of the catalyst. The surface-exposed active metal Ni and the strong acidity of the carrier surface favored the activation of the CO group of furfural. Meanwhile, the strong interaction between Ni particles and Fe3O4 carriers enhanced the stability of the catalyst. Wang et al. prepared catalysts of bimetallic Ni–Cu alloy loaded on ZSM-5 zeolite by a simple impregnation method. Through previous studies, it was demonstrated that the Cu/ZSM-5 catalyst was able to modify the adsorption mode of the CO bond in furfural, avoid saturation of the furan ring by presenting a tilted adsorption, but the conversion was low. While adding Ni to Cu/ZSM-5 in appropriate amount can improve the conversion and enhance the adsorption of furfural and hydrogen molecules at the same time. In addition, the acidity of the zeolite surface favors the cleavage of the C–O bond in the intermediate FOL. After 30 min of reaction at 220 °C, the yield of 2-MF was 78.8%.157 Gong et al. synthesized a nickel–cobalt bimetallic catalyst (NiCo–MgAlO) catalyst using a co-precipitation method, achieving 100% furfural conversion and 92.3% 2-MF selectivity.140 Bimetallic Ni–Co–MgAlO catalysts with uniform Ni–Co alloy nanoparticle compositions and more medium acid sites than single-metal catalysts are able to promote C–O bond breaking to improve 2-MF selectivity due to their oxytropism. Dou et al. prepared a CuCo alloy bimetallic catalyst (CuCo/NC) packed in a hollow nitrogen-doped carbon cage using ZIF-67 as a sacrificial template. After characterization and DFT calculations, it was shown that introduced Cu species, on the one hand, modified the electronic structure of Co, leading to a change in the adsorption configuration of furfural from parallel to vertical on the surface of the catalyst, which successfully impeded the hydrogenation of the furan ring. On the other hand, the formation of CuCo(111) crystalline face promotes hydrogen dissociation, as well as the cleavage of the C–O single bond, which lowers the diffusion barrier of hydrogen. The two aspects together promote the formation of 2-MF. Under relatively mild reaction conditions (150 °C, 1.5 MPa H2 and 4 h), the yield of 2-MF was as high as 95.7%.158 An et al. investigated transition metal oxide-loaded Cu catalysts prepared by sol–gel method with Mo as an additive (Cu–Mo/CoOx) to catalyze the hydrodeoxygenation of furfural to produce 2-MF. At Cu/Mo molar ratio of 3, furfural conversion exceeded 99% and 2-MF yield was as high as 92% at 180 °C and 2 MPa H2.144 XRD, XPS and CO-DRIFTS showed that the addition of Mo not only accelerated the formation of CuCo alloys in the catalysts, but also due to the electron transfer from Co to Mo. Enhanced the positive electronegativity of Coδ+, which in turn affected the adsorption of aldehyde groups on the carriers. The abundant Lewis acid sites on CoOx promoted C–O cleavage, leading to format more 2-MF. Koley et al. prepared Cu-based/mixed metal oxide catalysts (Cu/CuFe2O4@C-A) loaded on carbon using metal–organic skeletons as sacrificial templates. In this case, Cu can help furfural hydrogenation to alcohols, while the metal oxides can contribute to the hydroxyl removal step in the hydrodeoxygenation process.159 In Cu/CuFe2O4@C-A catalyst, the charge transfer between the carbon substrate and the metal oxides promotes the stabilization of the active metal sites in the ordered carbon framework, which contributes to the catalytic efficiency. About 90% yield of 2-MF was obtained by reacting at 165 °C, 1.5 MPa H2 for 5 h.
For preparation of 2-MF by catalytic hydrogenation and dehydration of furfural, liquid-phase hydrogenation is usually prone to side reactions and has lower selectivity. Niu et al. using isopropanol as a hydrogen donor, produced 2-MF by catalytic transfer hydrogenation of furfural in the liquid phase via CuZnAl catalyst. The results show that the intermediate FOL can be generated immediately in the reaction system. Subsequent C–O cleavage is the decisive step in the overall reaction. This step depends on the amount of Cu in the catalyst. The synergistic effect of Cu0 and Cu+ species on the catalyst surface, as well as the CuAl2O4 phase, gives the catalyst a good catalytic performance. For 4 h of reaction at 180 °C, the 2-MF yield was 72%.160 More et al. investigated MFe2O4 (M = Cu2+, Ni2+, Fe2+) antispinel catalysts and prepared several kinds of catalysts with magnetic properties and easy separation. Among them, CuFe2O4 was able to achieve 99.4% furfural conversion and 97.6% 2-MF selectivity at 200 °C over 1.5 h. In contrast, NiFe2O4 and Fe3O4 were mainly producing the hydrogenation product FOL. Upon characterization and analysis, it was found that the acidity of the catalyst is derived from the Lewis acid site, and its strength is directly related to the binding energy of Fen+ ions in the MFe2O4. The Cu2+ in CuFe2O4 has a stronger interaction with isopropanol, while the Fe3+ acts as active Lewis acid site centers, both of which work together to improve the catalytic activity of the catalyst.
Fig. 9 Plausible reaction mechanisms of (a) furfural hydrodeoxygenation to 2-MF on the surface of NiMo IMC, (b) furfural to 2-MF over CuCo/NC, (c) furfural to 2-MF over Co/CoAl2O4 catalyst, (d) copper oxides catalyzed furfural transfer hydrogenation in methanol.155,158,161,163 |
Dou et al. similarly deduced the reaction pathway for this reaction by theoretical calculations (Fig. 9b):158 H2 initially adsorbs on the active sites of Co and CuCo alloy and is dissociated into activated hydrogen. The carbonyl group of furfural molecule is firstly adsorbed vertically on the CuCo(111) interface and reacts with activated H atoms to form the intermediate FOL. Subsequently, the hydrodeoxygenation of FOL was realized due to the strong destructive effect of CuCo alloy on the C–O bond of FOL and the strong spreading effect of H species to obtain the final product 2-MF. Eventually, 2-MF was desorbed from the catalyst and the whole hydrodeoxygenation reaction process was successfully completed.
Li et al. then proposed a new concept – the important role of oxygen vacancies during the reaction (Fig. 9c).161 This is a very interesting and novel statement. They pointed out that the mechanism of the reaction could be: first, the oxygen vacancies enhance the adsorption of furfural by CoAl2O4, resulting in vertical adsorption of furfural on carrier. In this case, the O atom of the aldehyde group inserts into the oxygen vacancy, while the O atom of the furfuryl ring adsorbs on the Co active site. Subsequently, the activated hydrogen dissociates from Co0 and combines with the aldehyde group to form FOL, and then FOL desorbs from the catalyst. For FOL, the presence of oxygen vacancies does not change its adsorption configuration on the surface, but greatly increases its adsorption capacity. FOL sequentially undergoes C–O cleavage, hydroxyl hydrogenation, water desorption, and methylene furan hydrogenation to obtain 2-MF. By theoretical calculations, they suggest that the oxygen vacancies appropriately weaken the adsorption of activated hydrogen by Co in the neighboring active site, thus promoting the spillover of Co-desorbed activated hydrogen to FOL, and thus lowering the energy barrier from FOL to 2-MF.
Zhang et al. focused on transfer hydrogenation catalytic mechanism with methanol as the hydrogen donor (Fig. 9d).163 First, methanol adsorption reforming occurs on CuO(200)/Cu2O(111), splits methanol into CO and H2. Due to the presence of H2, Cu2+/Cu+ is partially reduced to Cu0 at 200 °C. In addition, a large amount of CuOx remains in the catalyst, providing an indispensable acid site for furfural conversion. Subsequently, H atoms adsorbed at Cu0 attack the CO bond to obtain the intermediate product FOL. Then, the O atoms on the FOL hydroxyl group are activated by Brønsted acid site to form transition state E. And then the reactive H atoms attack the α-positions of O and C atoms to form transition state F with the help of the Brønsted acid site, and be cleaved into a methyl group and a H2O molecule. The hydrogenolysis product 2-MF was finally obtained.
Overall, the adjustment of the adsorption configuration of furfural molecules on the catalyst surface and the coordination of Lewis acid and Brønsted acid sites on the catalyst carrier surface is very important for both gas-phase hydrogenation and transfer hydrogenation. In general, the furfural molecules should be controlled to maintain a tilted or even vertical adsorption on the surface, and at the same time, the adsorption on the CO bond should be strengthened as much as possible. This will benefit the generation of the intermediate FOL and effectively inhibit the saturated hydrogenation of the furan ring. The investigators’ study of the reaction mechanism for the preparation of 2-MF from furfural plays an important role in the subsequent directed design of catalysts and in the effective modulation of product selectivity.
With the continuous development of the economy, people's demand for PeDs is growing rapidly, but in recent decades, the traditional PeDs preparation is still mostly produced from petrochemical sources, which does not meet the green development concept, and there are widespread shortcomings such as serious equipment rot, high production costs and complex processes, which greatly restricts the large-scale production of PeDs.164 Due to the limited production capacity, usually in short supply, so the efficient synthesis of PeDs has excellent development prospects. Compared with the route using fossil energy as raw material, the preparation of 1,5-PeD using biomass-based furfural as raw material requires fewer reaction steps, can avoid the use of toxic reagents, and has high atomic economy. However, because the synthesis technology is not mature and the product yield is low, how to develop a synthesis process route with low cost, mild reaction conditions and environmental friendliness is the key to breaking through the current PeDs production bottleneck and improving the PeDs yield, which has very important research value.165
Entry | Catalysts | Solvent | T/°C | t/h | H2 pressure/MPa | PeDs yield/% | Ref. | |
---|---|---|---|---|---|---|---|---|
1,5-PeD | 1,2-PeD | |||||||
1 | Pt/Co2AlO4 | — | 130 | 24 | 1.5 | 30.6 | 10.3 | 166 |
2 | Li-modified Pt/Co2AlO4 | — | 140 | 24 | 1.5 | 34.9 | 16.2 | 166 |
3 | Pt/HT | 2-PrOH | 150 | 8 | 1.0 | 8.0 | 73.0 | 167 |
4 | Pt@Co2+ | EtOH | 150 | 4 | 3.0 | 47.0 | 23.0 | 168 |
5 | Pt@Al2O3 | H2O | 45 | 8 | (NaBH4) | 75.2 | — | 169 |
6 | Pd/MMT-K10 | 2-PrOH | 220 | 5 | 3.5 | — | 66.0 | 170 |
7 | Pd–Ir–ReOx/SiO2 | H2O | Step 1: 40 | Step 1: 8 | 6.0 | 71.4 | 1.4 | 171 |
Step 2: 100 | Step 2: 72 | |||||||
8 | Ru–Ir–ReOx/SiO2 | H2O | Step 1: 40 | Step 1: 8 | 6.0 | 78.2 | 2.8 | 172 |
Step 2: 100 | Step 2: 32 | |||||||
9 | Rh/OMS-2 | MeOH | 160 | 8 | 3.0 | — | 87.0 | 173 |
10 | Ru–Sn/ZnO | 2-PrOH | 150 | 6 | 3.0 | — | 84.2 | 174 |
Date et al. prepared Pd/MMT-K10 metallic acid two-center catalyst with montmorillonite as the carrier and used it for furfural hydrogenation to prepare 1,2-PeD. The yield of 1,2-PeD was 66% in 2-PrOH at 220 °C and 3.5 MPa H2 for 5 h. Py-IR confirmed the presence of Brønsted acid in the catalyst, which leads to the cleavage of furan rings after the formation of the first intermediate of furfural hydrogenation.170 Liu et al. studied the one-pot selective conversion of furfural to 1,5-PeD by two-step reaction on Pd-added Ir–ReOx/SiO2 catalyst. With water as solvent, the reaction was first performed in 6 MPa H2 at 40 °C for 8 h, and then at 100 °C for 72 h under the same hydrogen pressure, the highest yield of 1,5-PeD was 71.4%.171
Noble metal Rh also has good hydrogenation activity, so Liu et al. introduced Rh on Ir–ReOx/SiO2 catalyst. Similar to the previous study, the highest yield of 1,5-PeD was 78.2%, under the hydrogen pressure of 6 MPa, first at 40 °C for 8 h, and then at 100 °C for 32 h, with water as solvent. The catalyst characterization results showed that the Rh–Ir–ReOx/SiO2 catalyst exhibits an Ir–Rh alloy particle structure partially covered by ReOx substance, and the formation of Ir–Rh alloy is a necessary condition for high hydrogenation activity.172 Pisal et al. synthesized a bifunctional catalyst (Rh/OMS-2) with base and metal sites by hydrothermal method. The yield of 1,2-PeD was 87.0% at 160 °C and 3 MPa H2 for 8 h. Based on the obtained results and the characterization of the catalyst, it is found that there is a metal–base synergistic effect between the catalyst base carrier and the active metal, and the highly selective production of 1,2-PeD can be achieved through the adsorption and dissociation of H2 at the metal site and the adsorption type of the substrate at the alkaline site. Ru-based catalysts are also commonly used in the study of PeDs synthesis of furfural.173 Upare et al. used Ru3Sn7 alloy-supported ZnO catalyst (Ru–Sn/ZnO) for catalytic conversion of furfural to 1,2-PeD, which can effectively limit the formation of 1,5-PeD and other hydrolytic byproducts. Furfural could be completely transformed and the yield of 1,2-PeD reached 84.2% in 2-PrOH at 150 °C and 3 MPa H2 for 6 h. The highly selective formation of 1,2-PeD in furfural is attributed to the synergistic effect of Ru3Sn7 alloy phase and SnOx on basic ZnO support.174
For research use, in addition to the commonly used stainless steel batch reactors, fixed-bed reactors have also been used for catalyst performance testing. Fixed-bed reactors refer to filling particle solid catalysts or reactants in the reactor, forming a certain height of the stacked bed. And then the gas or liquid materials pass the gap between particles through the fixed bed layer at the same time, to realize the heterogeneous reaction process. Fixed-bed reactors have a wide range of applications in industry due to the simple structure, high reaction efficiency, full utilization of catalysts and controllability. Brentzel et al. prepared Ru–TiO2 catalyst and developed a process to synthesize 1,5-PeD from furfural using dehydration/hydration, ring-opening tautomerization and hydrogenation reaction. With water as solvent, 1,5-PeD with 84% yield can be obtained at 110 °C, 2 MPa H2 and H2 flow rate of 40 mL min−1.175
Entry | Catalysts | Solvent | T/°C | t/h | H2 pressure/MPa | PeDs yield/% | Ref. | |
---|---|---|---|---|---|---|---|---|
1,5-PeD | 1,2-PeD | |||||||
1 | Ni–Y2O3 | 2-PrOH | 150 | 72 | 2.0 | 46.0 | 1.9 | 176 |
2 | Ni–La(OH)3 | 2-PrOH | 150 | 72 | 2.0 | 55.8 | 2.8 | 176 |
3 | Ni/WxC/SiO2 | 2-PrOH | 200 | 6 | 6.5 | — | 21.0 | 177 |
4 | Ni1.6Fe0.4MgAl | EtOH | 170 | 3 | 4.0 | 31.0 | 25.9 | 178 |
5 | Ni–CoOx–Al2O3 | EtOH | 160 | 6 | 3.0 | 47.5 | — | 179 |
6 | Cu1.8Mg1.2Al | 2-PrOH | 140 | 10 | 6.0 | 15.6 | 46.0 | 180 |
7 | CuxCo3−xAl | EtOH | 140 | 6 | 4.0 | 41.1 | 10.0 | 182 |
Wijaya et al. prepared Ni–Y2O3 catalyzed hydrogenolysis of furfural by coprecipitation method. 2-PrOH was used as solvent and at 150 °C, 2 MPa H2 for 72 h, the yield of 1,5-PeD was 46.0% and that of 1,2-PeD was 1.9%, showing good selectivity of 1,5-PeD. When changing the support, Ni–La(OH)3 catalyzed furfural hydrogenolysis under the same conditions, 1,5-PeD yield increased to 55.8%, 1,2-PeD yield was 2.8%, also maintained a good 1,5-PeD selectivity. It is proved by characterization that Y3+ or La3+ has strong interaction with furfural carbonyl, but weak interaction with oxygen in furan ring, which is conducive to the selective breaking of C–O bond.176 Bretzler et al. used Ni/WxC/SiO2 in 2-PrOH, at 200 °C, 6.5 MPa H2 for 6 h, the combination of Ni and WxC in the catalyst to open the furan ring to produce 1,2-PeD, the final yield of 21%.177 Shao et al. synthesized bimetallic Ni–Fe catalysts (Ni1.6Fe0.4MgAl). The results show that the lower Fe content of catalyst contributes to the formation of mesoporous structure, but the higher Fe content destroys the LDH structure and leads to the decrease of surface area. Fe can also react with the metal Ni to form Ni–Fe alloys, which reduces the catalytic activity of further hydrogenation of furan rings in FA, so furfural will have the opportunity to produce PeDs through hydrogenolysis reaction. In ethanol, at 170 °C, 4 MPa H2 for 3 h, the yield of 1,5-PeD was 31%, and that of 1,2-PeD was 25.9%.178 Kurniawan et al. prepared Ni–CoOx–Al2O3 with the same hydrotalcite structure, and in ethanol at 160 °C and 3 MPa H2 for 6 h, the yield of 1,5-PeD reached 47.5%. Ni0 promotes the dissociation and adsorption of H2 molecules, and the liberated H atoms are then transferred to the oxygen vacancy CoOx (OV–CoOx) site through the hydrogen overflow mechanism. Then the H atom attacks the CO group, and finally the furan ring is cleaved to form 1,5-PeD.179 Shao et al. synthesized a layered bimetallic oxide Cu1.8Mg1.2Al catalyst with highly dispersed Cu particles and adjustable base site, and at 140 °C and 6 MPa H2 for 10 h to obtain a 1,2-PeD yield of 46%. The different content of Mg in the catalyst promoted the dispersion of copper oxide and the exposure of metal copper, weakened the interaction between copper oxide and the support, inhibited the agglomeration of metal copper substances, and promoted the catalytic activity.180 Fu et al. prepared CuMgAlO catalyst by urea decomposition method and at 150 °C, 6 MPa H2 for 6 h, the furfural conversion was 84.1%, and the selectivity of 1,2-PeD and 1,5-PeD was 54.9% and 27.8%, respectively. Through the characterization results, it was found that the ratio of Cu+/Cu2+ and the amount of acid and base played a key role in the catalytic performance of the catalyst during the calcination process, indicating that the selectivity of PeDs depends on the synergistic effect of Cu+ and acid and base properties.181 Hai et al. synthesized a series of CuxCo3−xAl hydrotalcite based catalysts with different Cu/Co molar ratios. When the Cu/Co molar ratio was 1:29 (Cu0.1Co2.9Al), the furfural conversion was 100% and the total PeDs yield was 51.1% after reaction in ethanol at 140 °C and 4 MPa H2 for 6 h. The yield of 1,5-PeD was 41.1%.182
The preparation of PeDs by furfural catalyzed conversion has also been studied using non-noble metal-based catalysts in fixed-bed reactors. Gavilà et al. attempted to convert furfural directly to 1,5-PeD in a spinel-structured Co–Al catalyst.183 After the reduction of spinel at 500 °C, the selectivity of FOL was high. However, at a higher reduction temperature (700 °C), furfural can be catalyzed to directly generate PeDs. The yield of 1,5-PeD was 30% after 8 h with 2-PrOH as the solvent at 150 °C and 3 MPa H2. The difference in catalyst activity at different reduction temperatures was attributed to the formation of cobalt nanoparticles in different states on the surface after high temperature reduction.
Date et al. investigated the reaction mechanism of furfural on the catalyst Pd/MMT-K10 to generate 1,2-PeD (Fig. 11a). It was deduced experimentally that the reaction started with the hydrogenation of the carbonyl group of furfural at the Pd site to produce FOL, followed by the extraction of H+ from MMT-K10 via the O of the furan ring (I) and simultaneous semi-hydrogenation of the ring to give the intermediate shown in (II), which is adsorbed on the surface of the catalyst via the C5–Pd bond. Such an adsorbed conformation allowed the subsequent hydrolytic breaking of the C5–O1 bond, followed by the formation of 1,2-PeD.170 Kurniawan et al. explored the mechanism of furfural conversion to 1,5-PeD on N4.5C3.5A1(1:2)-R (Fig. 11b). First, the H2 molecule adsorbed on the Ni0 site dissociates into H atoms overflowing into the CoOx site. Subsequently, the O1 atom in the furfural carbonyl group adsorbs on the Coδ+ site. The two H atoms then attack the C–O group of the adsorbed furfural to produce FOL, while the OV–CoOx on the catalyst surface will preferentially induce the breaking of the C2–O bond. After ring opening, H attacks the double bond and O2 atom in the furan ring to produce 1,5-PeD.179 The mechanisms of furfural ring-opening to produce PeDs are various so far, but it has been proved that the hydrogen overflow effect and the presence of oxygen vacancies are favorable for the reaction, so the design and preparation of catalysts on this basis will be beneficial to further improve the selectivity of PeDs.
Fig. 11 Proposed reaction mechanism for the direct conversion of (a) furfural to 1,2-PeD over Pd/MMT-K 10, (b) furfural to 1,5-PeD over N4.5C3.5A1(1:2)-R.170,179 |
In general, the direct conversion of furfural to PeDs is difficult and the yield is relatively low. This is because the adsorption type of furfural on the catalyst surface is related to many factors, and the reaction course may be conducive to the occurrence of other hydrogenation or hydrogenolysis reactions, which is not conducive to further ring opening. There are still some shortcomings in the current research, such as high cost of noble metal catalysts, harsh reaction conditions of non-noble metals, low product yield, poor stability, and unclear catalytic reaction mechanism. Although some research progress has been made compared to the past, there are still several problems to be solved in converting furfural directly to PeDs: (1) the specific action mechanism of mixed-valent oxide species in the reaction state, as well as related experiments and theoretical verification. (2) The non-noble metal-based catalyst used to avoid high-cost waste of noble metals has a relatively low yield of direct conversion of furfural to PeDs with high selectivity. (3) The selective ring-opening of furan rings in furfural is highly dependent on the adsorption types of furan rings and their intermediates as well as the order of hydrogenation and ring-opening.
Some studies showed that, in organic solvents, furfural hydrogenation is easy to form FOL, THFA, 2-MF, 2-THMF and other by-products, and furfural preparation of CPO reaction, water is essential, and weakly acidic or neutral aqueous solution is conducive to the generation of CPO.188–190 The following is a review and analysis of the reactions of furfural hydrogenation to CPO catalyzed by metal-based catalysts in aqueous media in recent years (Table 10).
Entry | Catalysts | T/°C | t/h | H2 pressure/MPa | CPO yield/% | Ref. |
---|---|---|---|---|---|---|
1 | Pd/UiO-66-NO2 | 150 | 5 | 1.0 | 95.8 | 191 |
2 | Pd/7.74% Y2(Sn0.65Al0.35)2O7−δ/Al2O3 | 150 | 6 | 4.0 | 98.1 | 192 |
3 | Pd/La2Ti2O7 | 150 | 6 | 4.0 | >82 | 194 |
4 | Pd/Cu-BTC | 150 | 24 | 4.0 | >90 | 198 |
5 | Pd/Fe-MIL-100 | 150 | 6 | 4.0 | 93.0 | 199 |
6 | Pd/FeZn-DMC | 150 | 6 | 4.0 | 96.0 | 200 |
7 | Pd–Co@UiO-66 | 120 | 12 | 3.0 | 95.0 | 201 |
8 | PdZn/ZnO | 120 | 12 | 4.0 | >90 | 202 |
9 | Pt/NC-BS-800 | 150 | 4 | 3.0 | >76 | 196 |
10 | Au/TiO2 | 160 | 15 | 4.0 | >99 | 197 |
11 | Ru/MIL-101 | 160 | 2.5 | 4.0 | >96 | 190 |
Liu et al. used porous carbon doped with heteroatoms as the support to prepare Pt/C catalyst. Due to its graded porous structure, high content of nitrogen and oxygen functional groups, highly dispersed Pt nanoparticles, good water dispersibility and reaction stability, it showed good activity for the hydrogenation of furfural. When Pt/NC-BS-800 is used as catalyst, the CPO yield was >76% at 150 °C and 3 MPa H2.196
Au-based catalyst has a weak catalytic hydrogenation ability, it is suitable for the conversion of furfural to CPO without producing other side reactions such as furan ring hydrogenation. Zhang et al. used anatase TiO2 nanorods with only mild Lewis acidic sites as carriers to load Au. At 160 °C, 4 MPa H2 pressure for 15 h, the conversion of furfural was 99%, and the CPO yield was >99%. The combination of Ru and acid center is also the preferred catalytic system for the preparation of CPO from furfural.197 Fang et al. designed a supported Ru nanoparticle catalyst (Ru/MIL-101) that achieved complete conversion of furfural within 2.5 h at 160 °C and 4.0 MPa H2, with a selectivity of >96%.190
The bimetallic supported catalyst also showed high activity in the catalytic hydrogenation of furfural to prepare CPO. The strong bimetallic synergistic effect between the two metal components was used to improve the catalytic performance of the catalyst. Deng et al. synthesized a Pd/Cu-BTC catalyst with Lewis acidity. Due to the high dispersion of Pd nanoparticles, furfural was almost completely transformed and CPO was more than 90% selective at 150 °C and 4.0 MPa H2 for 24 h.198 Li et al. found that with Pd/Fe-MIL-100 as catalyst, furfural conversion was 99%, and CPO selectivity was 93.1% at 150 °C, 4.0 MPa H2 for 6 h. When catalyst was reused for 5 times, its catalytic activity did not decrease.199 They also found that, using Pd/FeZn-DMC as catalyst, the furfural conversion was 99.9% and the CPO selectivity was 96.6% under the condition of 150 °C, 4.0 MPa H2 for 6 h. The catalyst was reused for 6 times, and the catalytic performance was almost unchanged.200 Wang et al. synthesized a core–shell Pd–Co@UiO-66 catalyst with furfural conversion of 99% and CPO selectivity of 96% at 120 °C and 3 MPa H2 for 12 h.201 Zhang et al. used PdZn/ZnO as catalyst at 120 °C and 4.0 MPa H2 for 12 h, and found that the catalyst showed catalytic generality for the synthesis of CPO from furfural compounds (i.e., furfural, 5-hydroxymethylfurfural and 5-ethyl furfural) and furanone (i.e., 2-furanmethyl ketone and 2-furanethyl ketone), the yield was over 90%.202
Entry | Catalysts | T/°C | t/h | H2 pressure/MPa | CPO yield/% | Ref. |
---|---|---|---|---|---|---|
1 | Cu/ZnO | 150 | 4 | 1.5 | 91.3 | 203 |
2 | Ni/SiO2 | 160 | 3 | 3.0 | 83.5 | 204 |
3 | Ni–P/γ-Al2O3 | 150 | 2 | 3.0 | 85.5 | 205 |
4 | Co@NCNTs | 140 | 5 | 4.0 | 75.3 | 206 |
5 | CoNP@N-CNTs | 160 | 8 | 0.5 | 95 | 207 |
6 | Cu0.4Mg5.6Al2 | 180 | 4 | 0.2 | 98.1 | 209 |
7 | Cu/Fe3O4 | 170 | 4 | 3.0 | 91 | 210 |
8 | CuNi/Al-MCM-41 | 160 | 5 | 2.0 | 97.0 | 211 |
9 | Ni2Cu1/Al2O3 | 140 | 1 | 1.0 | 89.5 | 212 |
10 | NiFe/SBA-15 | 160 | 6 | 3.4 | 90 | 189 |
The single metal-based catalyst showed good catalytic effect on the conversion of furfural to CPO. However, some inherent defects of single metals, such as the activity of Cu is usually low, requiring harsh reaction conditions, on the contrary, the highly active Ni catalyst is easy to lead to excessive hydrogenation of FOL to form THFA, limiting the further improvement of comprehensive catalytic performance. Therefore, combining different catalytic active centers to regulate the electronic structure through metal–metal interactions is a good way to achieve high-performance catalysts.
Zhou et al. prepared Cu0.4Mg5.6Al2 hydrotalcite catalyst by co-precipitation and calcination. The yield of CPO was 98.1%, the reaction temperature was 180 °C, the initial hydrogen pressure was 0.2 MPa, and the reaction time was 4 h.209 Pan et al. prepared magnetic Cu/Fe3O4 nanoparticles by coprecipitation. The maximum yield of CPO reached 91% at 3 MPa H2 and 170 °C for 4 h. The obvious interaction between the two active metals, Cu and Fe, is the main reason for the high selectivity of the catalyst for CPO.210 Zhang et al. synthesized the CuNi/Al-MCM-41 catalyst and reacted it at 2.0 MPa H2 and 160 °C for 5 h, obtaining 99.0% furfural conversion and 97.7% CPO selectivity. The small amount of aluminum highly dispersed in MCM-41 acts as an anchor and ensures the formation of highly dispersed CuNi bimetallic nanoparticles, which is conducive to the improvement of catalytic properties.211 Liu et al. synthesized a bimetallic NiCu nanoparticle catalyst with surface oxygen modification and reacted at 140 °C and 1 MPa H2 for 1 h, CPO yield of about 89.5%.212 Jia et al. used the bimetallic synergism between Ni and Fe in NiFe/SBA-15 catalyst to obtain a CPO yield of 90% at 160 °C, 3.4 MPa H2 for 6 h.189 Huang et al. prepared a multifunctional catalyst (xCo–yNi@NC) with highly dispersed Co–Ni alloy nanoparticles embedded in a porous carbon nitrided matrix by pyrolysis of a MOF template. Co–Ni alloy showed a strong bimetallic synergistic effect, and the introduced N species affected the physicochemical properties of the catalyst, thus promoting the catalytic performance. At 150 °C and 1.5 MPa H2 for 6 h, the CPO yield was 92.5%.213 Lin et al. synthesized the catalyst Ni3Sn2–ReOx/TiO2 by continuous impregnation method with 92.5% CPO selectivity at 140 °C and 3.0 MPa H2. Characterization demonstrated that the Ni3Sn2 phase significantly inhibited the overhydrogenation of the furan ring, whereas the Lewis acid site on ReOx activated the hydroxyl group of FOL and induced it to undergo rearrangement.214
Fig. 12 Reaction pathways for furfural to CPO.215 |
Mironenko et al. used D2O as a solvent for furfural hydrogenation over a 1% Pd/CNTs catalyst. It was found that furfural was hydrogenated to produce FOL, and then FOL was hydrogenated to produce the product CPO after Piancatelli rearrangement. The intermediate products of the reaction were detected as FOL, 3-hydroxycyclopentenone (HCPO), 4-hydroxy-2-cyclopentenone (HCPEO), etc. It was inferred that the reaction process involved hydrogenation, hydrolysis, intramolecular aldol condensation and hydroxylation. From this, the reactions were deduced that hydrogenation, hydrolysis, intramolecular aldehyde condensation, dehydration and hydrogenation reactions occurred during the reaction (Fig. 13).216 The reaction mechanism was then deduced as follows: selective hydrogenation of furfural to produce FOL, hydrolysis and dehydration of FOL to produce 2-pentene-1,4-dione, intramolecular aldol condensation to produce HCPEO, final hydrogenation to produce HCPO, and final hydrogenation and dehydration to produce CPO.
Fig. 13 Reaction network for the aqueous-phase hydrogenation of FAL over Pd/C catalysts. Undetectable intermediates are enclosed in square brackets.216 |
In the catalytic hydrogenation of furfural to prepare CPO, the metal active site and acid center on the catalyst are the key points of the catalytic reaction, and the temperature, pressure and reaction time of the reaction system all affect the selectivity of CPO. Currently, the gap between production and demand of CPO is large, and the preparation of CPO from biomass-based furfural is a green and environmentally friendly new technology in line with the development trend of atomic economy. This technology has very significant economic and social value. Finding and developing efficient catalytic systems for selective conversion of furfural to CPO remains a focus of future research. Non-noble metals are cheaper and easier to obtain than precious metals, and the catalytic effect of non-noble metal catalysts is better. Cu has the characteristics of high activity and selectivity in the catalytic reaction, and Ni also has better catalytic hydrogenation performance. However, due to the poor stability of catalyst and poor reuse, the research on non-noble metal-based catalyst is mainly to improve its stability and reusability. Bimetal catalyst has better catalytic performance than single catalyst, and the interaction between bimetals can further promote furfural hydrosynthesis of CPO, which has a wider application prospect.
Entry | Catalysts | Reaction conditions | FAM yield/% | Ref. |
---|---|---|---|---|
1 | Ru-PVP/HAP | 1 mmol furfural, 100 mg catalyst, 3 mL NH3 aq. solution, 100 °C, 0.4 MPa H2, 2 h | 60.0 | 223 |
2 | 2% Ru/MMT | 0.5 mmol furfural, 30 mg catalyst, 4 mL NH3 aq. solution, 90 °C, 1 MPa H2, 3 h | 89.0 | 226 |
3 | 1% Ru/TiO2 | 0.1 mL furfural, 25 mg catalyst, 10 mL NH3 aq. solution, 120 °C, 2 MPa H2, 2 h | 99.0 | 238 |
4 | Rh/Al2O3 | 0.2 g furfural, 2 mg catalyst, substrate/ammonia = 0.03, 80 °C, 2 MPa H2, 2 h | 91.5 | 231 |
5 | 12.5% Ni/MMT | 0.5 mmol furfural, 30 mg catalyst, 4 mL NH3 aq. solution, 130 °C, 1.5 MPa H2, 7 h | 84.0 | 226 |
6 | Ni/CaCO3 | 5 mmol furfural, 100 mg catalyst, 0.68 g NH3 aq. solution, 15 mL H2O, 7 mmol Zn, 80 °C, 10 h | 91.0 | 239 |
Chandra et al. prepared face centered cubic (FCC) ruthenium nanoparticles (Ru-NP) with a flat shape, and the selectivity for FAM could reach 99% with furfural as the substrate and methanol as the solvent with the addition of 8 mmol of NH3 at 90 °C, 2 MPa H2 for 6 h. The selectivity for FAM could reach 99% with the use of furfural as the substrate and methanol as the solvent. The catalyst exhibited the highest turnover frequency (TOF) of about 1850 h−1 during furfural reduction amination, the high performance of the catalyst was attributed to most of the metal Ru active sites with weak electron-donating ability.222 Nishimura et al. used poly(N-vinyl-2-pyrrolidone)-coated ruthenium-loaded hydroxyapatite (Ru-PVP/HAP) catalysts to show significant activity in the synthesis of FAM from furfural. The FAM yield was close to 60% when 5 wt% Ru-PVP/HAP was reacted in 25% NH3 aqueous solution for 2 h at 100 °C and 0.4 MPa H2. Meanwhile, Nishimura et al. demonstrated once again the importance of Ru metal as an active center by XAS spectral analysis.223 Deng et al. found that the effect of catalyst reduction temperature on the synthesis of FAM by low-temperature catalytic reductive amination of furfuryl aldehyde on Ru/Nb2O5·nH2O was significant due to the fact that the reduction temperature was directly related to the number of acidic sites on the catalyst. Ru/Nb2O5·nH2O possessed the highest density of acidic sites and the highest catalytic activity when reduced at 300 °C, and the reaction was carried out in methanol at 70 °C, 0.1 MPa NH3, and 3 MPa H2 for 6 h. The final yield of FAM could be obtained as 87%.224 Dong et al. showed that the strong interaction between Ru and Al in the carrier in the Ru/HZSM-5 catalyst increased the acid center sites on the catalyst and had a significant effect on the product distribution. The yield of FAM was 76% after Ru/HZSM-5 catalytic reduction of aminated furfural at 80 °C and 3 MPa H2 for 0.25 h. The FAM yield showed a decreasing trend when the reaction time was continued.225 Gokhale et al. loaded Ru on montmorillonite clay (MMT) in an aqueous solution of molecular hydrogen and liquid ammonia, the reductive amination of crude furfural extracted from rice husk to FAM could be achieved. The reaction was carried out in 2 mL of 25% NH3 aqueous solution at 130 °C, 0.25 MPa H2, for 7 h, with a FAM yield of up to 71%.226 Gao et al. using the characteristics of layered double hydroxide metal, as a carrier, the preparation of a kind of guarantee four ligand Al3+ (Al4c) and six coordination Al3+ loci (Al6c) as a complementary function, on the space close to each other and keep a certain distance with Ru metal site catalyst (Ru/Ni1MgAlOx). On Ru/Ni1MgAlOx catalyst, tetra-coordination and hexa-coordination trivalent aluminum can activate NH3 and nitrogen-containing compounds. The yield of FAM can reach 91.3% by adding 4.5 mmol NH3 to methanol at 2 MPa H2 at 90 °C for 5 h.227 Zou et al. used boron nitrogen co-doped carbon (BNC) as a carrier-supported Ru catalyst (Ru/BNC) with a surface rich in hindered Lewis acid–base pairs (FLPs) and hydrazine hydrate (N2H4·H2O) as nitrogen source for reducing amination reaction. Methanol as a solvent, at 0.24 mmol N2H4·H2O, 2.0 MPa H2, 80 °C for 16 h, the catalyst can efficiently and selectively promote the conversion of furfural to FAM, with a yield of more than 99%. Mechanistic studies showed that abundant FLPs on the BNC surface synergistically enhanced the activity of Ru catalyst. In addition, the rapid formation and moderate reactivity of the intermediate hydrazone significantly improved the selectivity of FAM.228 Wang et al. studied the catalytic activity and selectivity of various active metals in the furfural reduction amination reaction and found that Ru-based catalysts showed the highest catalytic activity compared with other metal-based catalysts. Reaction of 1 wt% Ru/α-Al2O3 catalyst in methanol, 0.2 MPa NH3, 2 MPa H2 at 70 °C for 24 h, the yield of FAM was 61.9%.229
In addition to Ru, noble metal-based catalysts such as Pd, Pt, and Rh are widely used in the catalytic reductive amination of carbonyl compounds due to the mild preparation conditions and simple methods. Wang et al. prepared Pd/MoO3−x catalysts and regulated the interaction between Pd nanoparticles and MoO3−x carriers by changing the preparation temperature. The reaction was carried out at 80 °C, 7 mmol NH3, 2 MPaH2 for 4 h. The yield of FAM could reach 84%. Low-valent Mo species in the catalysts can act as Lewis acid sites to activate the aldehyde group and provide strong metal–carrier interactions, enabling the metal Pd center to promote the amination of furfural.230 Chatterjee et al. used ammonia as the amine source and hydrogen as the reducing agent, the reductive amination reaction of furfural was carried out using Rh/Al2O3 catalyst at 80 °C for 2 h. The FAM selectivity could reach 91.5%.231
Ni was the first metal applied in aldehyde/ketone reductive amination reactions and is the most studied and reported active metal component of the catalyst. Zhou et al. directly used RANEY® nickel as a catalyst, tetrahydrofuran (THF) as a solvent, added 0.35 MPa NH3, 0.5 MPa H2, and reacted at 180 °C for 12 h. It was able to obtain FAM in 99.2% yield. DFT calculations showed that the difference in adsorption energies of NH3 and H2 on Ni is smaller than that of other metals, indicating that NH3 occupies fewer metal active sites on the Ni surface, which leaves more active sites for the dehydrogenation/hydrogenation reaction, and ultimately facilitates the reductive amination reaction.233 Song et al. prepared an efficient nitrogen-doped porous carbon-loaded nickel catalyst (Ni/pNC) by template-assisted pyrolytic impregnation, with a FAM yield of 92.3% in methanol solvent, 0.4 MPa NH3, 3 MPa H2, and 60 °C for 6 h. The characterization demonstrated that the reductive amination of furfural to FAM was associated with the formation of Ni–Nx sites on the surface of N-doped carbon and the electronic interactions between N and Ni, which significantly reduced the activation energy of the reductive amination of the reaction intermediate.234 Yang et al. loaded Ni nanoparticles uniformly on the surface of SiO2 carrier to prepare Ni/SiO2, the strong Lewis acidity is an important reason for its high reactivity and high selectivity. The yield of FAM could reach 98% at 0.8 MPa NH3, 4 MPa H2, and 90 °C.235
Co is another commonly used active metal component of catalysts among non-noble metals. Zhuang et al. prepared a graphene-co-shelled cobalt nanoparticles (Co@C-600), and the selectivity of FAM obtained by furfural reductive amination was able to reach 98.45% by reacting with 2.0 MPa H2 in 5 mL ammonia solution at 90 °C for 4 h. The reaction was carried out at 90 °C for 4 h. The selectivity of FAM obtained by furfural reductive amination was also able to reach 98.45%.236 Yogita et al. produced Co/NC-700 catalyst by pyrolysis using ZIF-67 as precursor, which had good recoverability while performing reductive amination reaction in methanol and liquid ammonia solution at 120 °C for 1 h at 2 MPa H2, and the yield of FAM reached 99% in a short time. The catalyst can also be used for direct one-pot conversion of xylose to FAM in 68.7% yield. The characterization results proved that the metal cobalt nanoparticles coordinated with nitrogen were the active centers for the reductive amination reaction. The presence of high specific surface area, easy hydrogen desorption and surface defect sites on the catalysts contribute to the catalytic activity and selectivity.237
Fig. 15 Three possible reaction pathways for the reductive amination of furfural to FAM. (a and b) With NH3 in the presence of H2 gas, (c) with the nitrogen source of N2H4·H2O.223,228,231 |
However, since secondary amines are formed by the interaction of primary amines with carbonyl compounds, the production of secondary amines and primary amines as by-products during the reaction process is unavoidable. The selectivity of FAM is also hampered by the formation of by-products originating mainly from hydrogenation and/or hydroxyaldol condensation. Therefore, in order to avoid the generation of hydrogenation products from furfural and to inhibit the condensation reaction, Zou et al. carried out a reductive amination reaction using N2H4·H2O as a nitrogen source. The mechanistic investigation showed that furfural was synthesized into (2-furanylmethylene)hydrazine and 1,2-bis(2-furanylmethylene)hydrazine under N2H4·H2O as nitrogen source, followed by hydrogenolysis to produce the product FAM under H2. The rapidity of the intermediate stilbenes in the reaction process formation of the intermediate hydrazone, as well as its moderate reactivity significantly improved the selectivity of FAM (Fig. 15c).228
Fig. 16 Reaction pathways for the reductive amination of furfural to (a) pyrrole, (b) piperidine, (c) indole. |
Piperidine is used as a condensing agent and solvent in organic synthesis, and can also be used in the manufacture of local anesthetics, painkillers, fungicides, wetting agents, epoxy resin curing agents, rubber vulcanization accelerators and so on.244 There are few studies on the direct one-step preparation of piperidines from furfural, a reaction cascade that starts with reductive amination of furfural, followed by hydrocyclic cleavage, dehydrogenation, intramolecular reductive amination and final hydrogenation to produce piperidines, especially the selective C–O cleavage of the furan ring under reductive amination conditions, which is the most key reaction step. Therefore, the key to realizing this new strategy lies in the fine design of multifunctional catalysts.245 In order to solve this challenge, Qi et al. targeted the design and synthesis of a Ru1CoNP/HAP catalyst (Ru species are sufficiently isolated by Co atoms and form surface single-atom Ru1CoNP alloy structure) which showed the highest yield of piperidine of 93% in the presence of NH3 and H2. Kinetic investigation verified the reaction pathway, initially, furfural was reductively aminated to produce FAM, then hydrogenated to tetrahydrofurfurylamine (THFAM), and after 6 h of reaction at 100 °C, and then warmed up to 180 °C for 14 h of reaction, the cyclic ether bond of THFAM was broken, and rearrangement to produce piperidine was initiated (Fig. 16b). DFT calculations demonstrated that the surface structure of Ru1CoNP facilitated the direct ring opening of THFAM, as well as the rapid conversion of the intermediate 5-amino-1-pentanol to the product piperidine. This study achieved a one-step conversion of furfural to piperidine for the first time.246
Indole is widely contained in lemon oil, citrus oil, grapefruit peel oil, jasmine oil and other essential oils, and can be widely used in jasmine, lilac, gardenia, lotus, narcissus and other floral flavors. Derivatives of indole are widely distributed in nature, and many natural compounds contain indole ring in their structure, and some indole derivatives are closely related to life activities, so indole is also a very important heterocyclic compound. The new efficient and sustainable synthetic method using the one-step conversion of furfural to indole offers the possibility of producing indole from renewable and environmentally friendly biomass resources. Yao et al. used HZSM-5 (Si/Al = 25) as catalyst and the maximum carbon yield of indole was 20.79% at 650 °C, an airspeed of 1.0 h−1, and the NH3 to furfural molar ratio of 2. Based on the experiments, Yao also verified a possible reaction pathway for the conversion of furfural to indole: furfural first reacts with ammonia to form imine, which subsequently undergoes a cleavage reaction to produce furan after decarbonization. After that, the furan continues to react with ammonia to form pyrrole, and is finally converted to indole by the Diels–Alder reaction, a dehydration reaction between furan and pyrrole (Fig. 16c).247
So far, noble metal catalysts can obtain high yields of primary amines under milder reaction conditions. The non-noble metal catalysts are inexpensive but have the problems of complicated preparation process, high loading and harsh reaction conditions. The particle size, nature, structure and specific surface area of the carriers have a significant influence on the catalytic performance of the catalysts. The heterogeneous catalysts for the catalytic reductive amination of carbonyl compounds to synthesize primary amines still have difficulties of low reusability of the catalysts due to the susceptibility of the metals to agglomeration, leaching of the metals, and poisoning of the metal active centers. There are also few reports on the generation of other N-heterocyclic compounds from biomass, but in fact nitrogen-containing compounds are usually of higher value compared to oxygenated chemicals and are widely used in the production of pharmaceuticals and agrochemicals.248 Therefore, the development of simple, efficient, highly selective and stable non-noble metal catalysts to catalyze the preparation of nitrogen-containing compounds from furfural is a major trend.
Acetone is often used as a substrate for the reaction of aldol condensation with furfural. A high concentration of moderately strong basic sites enables the successful condensation of furfural with acetone.251 Most of the industry uses aqueous sodium hydroxide and ammonia for the furfural condensation reaction with acetone. However, in order to have further development in terms of recoverability, separation and purification, heterogeneous solid base catalysts have been developed and investigated. Solid base catalysts in aldol condensation include basic or alkaline earth metal hydroxides, magnesium organosilicates, zeolites, alumina, and magnesium oxide (Table 13).252 Al–Zr mixed oxides have excellent activity for the acetalization reaction of furfural with acetone. But generally, mixed oxides may exhibit severe deactivation, which is associated with furfural deposition of compounds on the catalyst surface and/or leaching of active substances.253 Much work has been done to improve the stability of Al–Zr mixed oxides. Yuan et al. developed a simple low-temperature solution phase induction strategy for manufacturing Al–ZrO2 composites with tunable surface acid–base properties and high specific surface area. In the Al–ZrO2 composite, the catalyst with an Al/(Al + Zr) molar ratio of 0.5 had the highest activity, the furfural conversion was as high as 92.0% and the 4-(2-furyl)-3-buten-2-one (FAc) selectivity was 98.7%.254 The synergistic interaction of basic and acidic sites on the catalyst surface is crucial for the superiority of the catalyst in aldol condensation. Similar to Al–Zr oxides, mixed oxides such as Mg–Al have been developed as heterogeneous catalysts for the condensation of furfural and pyruvic aldehydes.255–257 Arhzaf et al. prepared a series of MgAl–CO3 type hydrotalcite materials, and the mixed oxide with a Mg/Al ratio of 3.5 (Mg3.5Al–O), prepared from hydrotalcite calcined at 450 °C, was the most alkaline and catalytically active of the series of catalysts.258 A molar ratio of furfural/acetone of 1/10 was added to the reactor and the reaction was carried out at 90 °C for 2 h. The conversion of furfural and the selectivity of FAc were optimized. The conversion was 98% and the selectivity to FAc was 78%. It realized a highly selective condensation reaction in the solvent-free condition.
Desai et al. made some studies on the reaction mechanism of the condensation of furfural with acetone (Fig. 17). The catalyst alkalinity they prepared varied with different molar ratios of Mg and La. When the Mg, La molar ratio was 3:1, the activity of the catalyst increased significantly due to the presence of medium to strong basic sites in the catalyst. At the beginning of the reaction, an acetone and a furfural adsorbed on the basic and acidic sites of the mixed oxide catalyst, respectively. Then, the enol ester formed on the surface of the catalyst attacks the carbonyl group of the furfural, which subsequently dehydrates to form FAc.259
Fig. 17 Reaction scheme for aldol condensation of furfural and acetone.259 |
Overall, increasing the distribution of basic sites can effectively improve the catalyst activity in the furfural condensation reaction with acetone.251 But most of the catalysts will be susceptible to deactivation due to the deposition of compounds on their surfaces during the reaction process. So, improving the stability of the catalysts, as well as optimizing the structure of the catalysts or/and adjusting the surface properties of the catalysts to address issues such as deactivation are of greater interest. As of now, the aldol condensation reactions of biomass-derived oxygenated furan compounds, such as furfural, have been less studied by heterogeneous acid catalysts compared to heterogeneous base catalysts. However, furan-containing compounds are usually produced from cellulose or hemicellulose under acidic conditions. To develop effective acidic catalysts for aldol condensation one-pot conversion from biomass furan oxide compounds, and hence eliminating neutralization and purification operations in the production process, is worth further exploration and research.
The high-value compounds into which furfural is converted can generally be divided into two categories: biomass-based chemicals and biofuels. When converting furfural into biofuel compounds, it is generally necessary to selectively remove oxygen atoms on carbonyl carbon to remove oxygen-containing functional groups and improve its octane number to meet the needs of fuel. The conversion of furfural into high-value biological chemicals is generally to retain oxygen atoms on carbonyl carbon, and the conversion of furfural into oxygen-containing compounds can make better use of this biomass-derived molecule with highly functional groups. In addition, by using furfural as a renewable platform compound, high-value compounds can be produced in fewer steps and are more environmentally friendly than by using fossil resources as raw materials, in line with the current green chemistry concept of sustainable production. This review systematically summarized the latest research progress in the efficient conversion of furfural to other high-value compounds, introduced a variety of efficient and green noble and non-noble metal catalytic systems, especially for catalytic reactions in green solvents, such as H2O. In multiphase catalytic systems, the conversion of furfural to value-added chemicals such as LA, GVL, PeDs and CPO is a very valuable study because these production pathways are highly atomistic-economic, sustainable, and very much in line with the requirements of green development.
However, the following key issues still need to be addressed in future research on the conversion of furfural to other value-added chemicals: (1) Catalyst performance: due to the complexity of the conversion pathway, and the competing side reactions, little research has been done on some high-value chemicals such as PeDs, pyrroles, and piperidines. Researchers are working to develop more efficient and selective catalysts for the targeted conversion of furfural to a variety of high-value products. These catalysts can control reaction conditions, improve product selectivity, and minimize by-product generation. (2) Catalyst design: most studies lack deeply rooted mechanisms and catalytic design strategies specifically explored and validated for the conversion of furfural to chemicals. Using theoretical modeling, kinetic studies and characterization tools, the role played by each active component and active site of the catalyst in the catalytic process can be verified, which facilitates the proposal of catalyst design strategies to further improve the reaction efficiency and target product selectivity. (3) Green catalytic technologies: green catalytic technologies aim to reduce energy consumption, minimize waste generation, and enhance the environmental friendliness of reactions. Therefore, the development of green catalytic technologies is crucial for furfural conversion processes, such as research into catalyst design based on renewable resources and the use of environmentally friendly solvents. (4) Exploration of novel catalytic mechanisms: furfural conversion to high-value compounds usually involves multiple reaction steps, but the synergistic effect between different metals of multimetallic composite catalysts and the adsorption configuration of the reaction substrate on the catalyst surface are not well understood and lack of corresponding theoretical validation. Confronted with problems such as the controversial reaction mechanism of some reactions, and because many target products become intermediates of other reactions under different conditions, how to control the selectivity of the target products becomes a major challenge. Therefore, when designing highly selective and active bifunctional catalysts, each step of the reaction should be matched with the corresponding active site and reaction conditions. (5) Expanding new downstream products: the multiple functional groups of furfural and its high reactivity make it a strong potential for development as a functionalized renewable platform compound. Besides the developed downstream products such as FOL, THFA, LA, GVL, PeDs, CPO, FAM, 2-MF, 2-THMF, etc., more value-added chemicals used in various fields can continue to be explored in the future to maximize the value of furfural.
In summary, prospects for catalytic technologies for downstream products of furfural include advancements in selective catalytic conversion, optimization of multiphase catalytic systems, catalyst regeneration and recycling, application of green catalytic technologies, exploration of novel catalytic mechanisms, and exploration of other downstream products. The continuous development of these technologies will drive the effective utilization of furfural and promote the development of sustainable chemical industries.
This journal is © The Royal Society of Chemistry 2024 |