Next Article in Journal
Exploring the Multifaceted Potential of 2D Bismuthene Multilayered Materials: From Synthesis to Environmental Applications and Future Directions
Next Article in Special Issue
VO Supported on Functionalized CNTs for Oxidative Conversion of Furfural to Maleic Anhydride
Previous Article in Journal
Editorial: A Special Issue on “Catalytic Processes in Biofuel Production and Biomass Valorization, 2nd Edition”
Previous Article in Special Issue
H-Beta Zeolite as Catalyst for the Conversion of Carbohydrates into 5-Hydroxymethylfurfural: The Role of Calcination Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Catalysts in the Conversion of Biomass and Its Derivatives

1
Xinjiang Biomass Solid Waste Resources Technology and Engineering Center, College of Chemistry and Environmental Science, Kashi University, Kashi 844000, China
2
College of Environmental Science and Engineering, Nankai University, Tianjin 300071, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(8), 499; https://doi.org/10.3390/catal14080499
Submission received: 30 June 2024 / Revised: 15 July 2024 / Accepted: 30 July 2024 / Published: 1 August 2024
(This article belongs to the Special Issue Catalytic Conversion of Biomass to Chemicals)

Abstract

:
With the continuous depletion of fossil resources and the deterioration of the global climate, it is particularly urgent to find green and sustainable renewable resources to replace non-renewable resources. Renewable biomass, which converts and stores light energy into chemical energy through photosynthesis by green plants, has received widespread attention due to its simultaneous resource and energy properties. Therefore, this article focuses on lignocellulose, an important component of biomass, in the fields of chemical conversion and high-value-added chemical preparation. A detailed review was conducted on the application of catalysts in biomass bio-char, bio-oil, bio-gas, and high-value added chemicals and their derivatives, represented by 5-hydroxymethylfurfural (5-HMF) and levulinic acid (LA). At the same time, the difficulties and challenges encountered by catalysts in biomass conversion were analyzed, and new ideas were proposed for future development directions, so as to provide new development pathways for efficient and green conversion of biomass into biomass energy and high-value-added chemicals.

1. Introduction

The extensive use of fossil resources such as coal, oil, and natural gas has greatly contributed to the rapid development of the modern economy and society and the continuous improvement of people’s living standards [1]. However, with the continuous increase in the world population, people’s demands for resources and energy consumption have also soared dramatically, and fossil resources, as non-renewable resources with limited reserves, have been unable to meet people’s needs for energy and resources, thus limiting economic and social development [2]. At the same time, in the past several hundred years of fossil resource extraction and use, subject to ideological and cultural education conditions, the level of scientific and technological development, economic and social development processes, and other factors, the ecological environment has caused non-negligible negative impacts, such as the greenhouse effect, global warming, and the destruction of the ozone layer, thus seriously affecting the sustainable development of people [3]. In addition, the uncertainty of geopolitical developments such as the Russian–Ukrainian conflict and the Palestinian–Israeli conflict, the prevalence of trade protectionism globally, and imbalances in the distribution of fossil resources across different regions have led to the unstable, uncontrollable, and unpredictable prices of fossil resources and energy sources, thus increasing the cost of living of people around the world and worries about future uncertainty. This has also slowed down and limited the pace of the global economic recovery [4]. Therefore, it is particularly important to find and develop renewable resources that can replace petrochemical resources.
Currently, various renewable resources have been developed, including biomass, wind energy, solar energy, and tidal energy [5]. However, among them, wind, solar, and tidal energy are intermittent energy sources, and their production is easily affected by weather conditions, thus increasing the uncertainty and instability of energy supply, especially under some extreme meteorological conditions [6]. In contrast, biomass, as the only renewable resource in the form of carbon-containing substances, has the advantages of abundant resources, wide distribution, and waste recycling, making it one of the most promising green alternative resources and renewable energy sources. [7].
Lignocellulose, as the most abundant renewable biomass resource on earth, mainly consists of cellulose (40–50 wt%), hemicellulose (15–30 wt%), and lignin (16–33 wt%) [8], as shown in Figure 1. Cellulose is a polymer made of glucose linked by hydrogen bonding and van der Waals forces, and its decomposition products are mainly low molecular compounds such as sugar aldehyde, formic acid, furan, CO2, and H2O [9]. Hemicellulose is a polysaccharide formed from pentoses (xylose, arabinose) and hexoses (glucose, galactose), etc., which can be used to produce ethanol and bio-gas [10]. Lignin has a complex three-dimensional network structure that tightly links cellulose and hemicellulose together and is commonly used to prepare aromatic compounds such as benzene, toluene, and xylene (BTX), phenol, and guaiacol [11].
Generally speaking, there are three main ways to utilize lignocellulose resources, as shown in Figure 2. The first direct utilization method is that the biomass is directly burned, producing heat. This method exists as a single form of energy utilization, has a low energy utilization efficiency, and is accompanied by environmental pollution problems [12,13]. The second is the bio-enzymatic method, which uses microorganisms or bio-enzymatic fermentation to convert biomass into bio-gas and bio-ethanol, which has a high production cost, a long conversion time, and a single product [14,15]. The third is the chemical conversion method, which is to convert biomass into various fuels and chemical products that can be easily stored and distributed by controlling various influencing factors in the reaction system. It has the characteristics of rapid production, rich product types, and a high resource utilization rate, and it has been widely used in the field of high-value-added biomass resource refining [16,17,18].
Chemical conversion of lignocellulose can produce three different types of products, namely biomass bio-char, bio-oil, and biomass combustible gas, and its further conversion can produce high-value-added chemicals (5-HMF and its derivatives, LA, methyl levulinate, γ-pentolactone, and 2,5-furan dicarboxylic acid), biofuels (ethanol, hydrogen, and methane), and novel biomass-based materials (activated carbon, adsorbents, and catalysts) [19,20], as shown in Figure 3.
Although the traditional chemical conversion of lignocellulose can achieve green and sustainable utilization of biomass resources, there are problems such as insufficient conversion of effective products, difficulties in separation and purification, and poor selectivity that seriously limit its large-scale production [21]. The introduction of catalysts in the chemical conversion process can effectively improve the selectivity, yield, and reaction rate of conversion products, as well as reduce the cost [22]. Therefore, this paper provides a detailed review of the application of catalysts in the conversion of biomass and its derivatives from both the chemical conversion products of lignocellulosic biomass and the production of high-value-added chemicals, as well as providing an outlook on the challenges in future development, so as to provide new ideas for the efficient and green conversion of biomass into sustainable biomass energy and high-value-added chemical products.

2. Application of Catalysts in Biomass Conversion

Biomass is mainly made up of converted inorganic substances, such as carbon dioxide and water, that become organic compounds rich in various functional groups through photosynthesis, which makes it possible to convert the light energy widely available in nature into renewable chemical energy that can be fully utilized, thus enabling it to show its potential to replace fossil fuels [23]. The chemical conversion of biomass is an effective way to achieve this transformation, especially by introducing catalysts into the field of biomass thermochemical conversion, which could convert biomass into various resources such as solid biomass bio-char, liquid bio-oil, and synthesis gas, thereby achieving fuller and more efficient utilization of biomass resources [24]. The effects of catalysts on biomass conversion products are shown in Table 1.

2.1. Application of Bio-Char Catalysts

Bio-char is a structurally dense and stable carbon-containing substrate formed during biomass conversion [37]. With different biomass raw materials and conversion conditions, bio-char can show different physical and chemical properties, which are widely used in many fields. For example, the high carbon content and developed pore structure of bio-char makes it a suitable soil amendment, which plays a role in carbon sequestration, improving soil fertility, and soil pollution remediation. The rich variety of functional groups and structurally stable organic carbon and minerals in biomass charcoal make it suitable for use as catalysts and their carriers, bio-energy storage materials, adsorbents, and other high-value sustainable platform carbon materials [38].
Liu et al. [39] prepared low-cost nitrogen-doped carbon materials with a graphene-like structure by the pyrolysis method in a tube furnace using waste lignin powder and Ni powder as raw materials and characterized their properties. The results showed that nitrogen-doped carbon materials synthesized from waste lignin had a large specific surface area (289.37 m2g−1) and exhibited excellent oxygen reduction reaction catalytic activity, good methanol tolerance, and remarkable long-term stability, and thus have potential application value in various fields such as in oxygen reduction reaction catalysts, catalyst carriers, and hydrogen fuel cells.
Kong et al. [40] prepared bio-char-based catalysts (Fe-MEC, Ni-MEC, and Co-MEC) loaded with nano-Fe, Ni, and Co from biomass pyrolysis using sequoia wood chips as raw material by a microwave heating method and used them for the catalytic pyrolysis of polystyrene powders. The results showed that the three bio-char-based catalysts prepared had a large specific surface area and a developed pore structure. The specific surface areas of Fe-MEC, Co-MEC, and Ni-MEC reached 307.31, 216.93, and 122.03 m2g−1, respectively. Meanwhile, the nanoparticles of Fe, Ni, and Co were also uniformly distributed on the surfaces of the carbon skeleton, which exhibited more catalytic active sites. Moreover, during the catalytic pyrolysis of polystyrene powder, the Fe, Ni, and Co nanoparticles loaded on the surface of the bio-char exhibited high catalytic activity for the breaking of C-C and C-H bonds in the bio-oil, which effectively increased the bio-oil yields (the bio-oil yields of the Fe-MEC, Ni-MEC, and Co-MEC catalysts were 91%, 82%, and 90%, respectively) and improved the quality of bio-oil. At the same time, Fe-MEC, Ni-MEC, and Co-MEC also showed excellent stability, and after recycling the catalyst five times, the yield of bio-oil did not change significantly.
Xu et al. [41] used coconut shell and ZnCl2 as templates and activators, respectively, to prepare biomass-based porous carbon carriers with large specific surface areas through pyrolysis in a tubular furnace and loaded ZnNiMoSx on them to prepare ZnNiMoSx/C composite catalysts. The results showed that ZnCl2 could effectively regulate the porous structure of the biomass carbon carriers to present a larger specific surface area (1581.1 m2g−1), thereby improving the dispersion of ZnNiMoSx on the biomass-based porous carbon carriers and promoting the formation of more active sites in the ZnNiMoSx/C composite catalysts to enhance their hydrodesulfurization activities. In addition, the ZnNiMoSx/C composite catalysts had a higher degree of sulfidation and more sulfur vacancies, thereby improving their catalytic efficiency for dibenzothiophene. The conversion rate was 99.7% and direct desulfurization selectivity was 94.5%, and stability could reach 40 h.
Lu et al. [42] used corncobs as biomass raw material combined with sodium alginate to prepare a bio-char-based alginate hydrogel catalyst and investigated its effect on phenol degradation. The results showed that the bio-char-based alginate hydrogel catalyst could effectively degrade phenol in wastewater, and its degradation efficiency could reach 69.2%. Meanwhile, the addition of bio-char materials promoted the activation of hydrogen peroxide, thus improving the ability of the bio-char-based alginate hydrogel to degrade pollutants, also showing excellent cycling stability. The degradation efficiency of phenol did not change significantly after three cycles of recycling.

2.2. Application of Catalysts in Bio-Oil

Bio-oil is an organic liquid produced during the process of biomass conversion and includes various components such as water, alcohols, acids, ketone, furan, hydrocarbons, nitrogen-containing compounds, and oxygen-containing compounds. It has the advantages of high energy density, convenient transportation, and rich functional groups. It has been used as a biomass fuel, organic solvent, and chemical raw material, thus demonstrating excellent resource recycling and environmental protection benefits, and it has become an effective way to replace fossil resources [43]. However, the bio-oil produced by traditional biomass conversion has the characteristics of a high oxygen content, poor stability, complex composition, and strong corrosiveness, which limit its high-value utilization [44,45]. Therefore, it is particularly important to find suitable methods to improve the quality and utilization value of bio-oil. In the process of biomass conversion, the addition of catalysts can not only reduce the activation energy of chemical reactions but also change the composition and selectivity of bio-oil, thereby obtaining high-quality biomass fuels and high-value-added chemicals [46].
Zhang et al. [47] used metal oxides such as Al2O3, SiO2, ZnO, MgO, and CaO as catalysts for the pyrolysis conversion of poplar, cellulose, and lignin in a fixed bed reactor. The results showed that the acidic sites in the acidic catalysts (Al2O3, SiO2, and ZnO) could effectively promote the cracking of macromolecular compounds and the dehydration reaction of organic molecules, inhibiting the further decomposition of volatiles into gaseous products in the pyrolysis conversion products, thereby improving the yield of bio-oil and also promote the formation of carboxylic acids, aldehydes/ketones, furans, and phenolics, while alkaline oxides (CaO and MgO) could significantly inhibit the formation of phenolic substances.
Li et al. [48] modified HZSM-5 with green templates (cellulose, starch, and glucose) to prepare a composite catalyst with a rich mesoporous structure and used them for the catalytic conversion of biomass (rice straw and rapeseed straw) to prepare aromatics in a fixed bed reactor. The results showed that the pore size of the template molecules had a significant impact on the catalytic activity of the composite catalyst, and the relatively large mesoporous structure promoted the selectivity of monocyclic aromatic hydrocarbons with higher carbon numbers, in which the composite catalysts prepared with cellulose and glucose as templates had greater selectivity for benzene, while the composite catalyst prepared with starch as a template had higher selectivity for toluene and xylene. Meanwhile, the composite catalysts had a larger specific surface area and more catalytically active sites than HZSM-5, which was more favorable for the production of aromatics and phenols, and the maximum yield of BTX was 10.18 wt%, which was 1.1 times that of HZSM-5.
Gupta et al. [49] used γ-Al2O3 and nickel-doped γ-Al2O3 (Ni/Al2O3) as catalysts for the catalytic conversion of waste pine needle biomass in a semi-intermittent reactor and investigated the pyrolysis characteristics and product fractionation. The results showed that the addition of catalysts facilitated reforming, cracking, and deoxygenation reactions, which led to a reduction in bio-oil yield and an increase in bio-gas yield. The addition of catalysts significantly reduced the content of oxygen-containing compounds such as ketones, furans, acids, sugars, and nitrogen-containing compounds in the pyrolysis products, while increasing the content of hydrocarbons and phenols in the bio-oil. The pyrolysis bio-oil with no catalyst, a γ-Al2O3 catalyst, and a Ni/Al2O3 catalyst contained 1.86%, 8.951%, and 20.25% hydrocarbons, as well as 15.31%, 20.13%, and 39.23% phenol contents, respectively. Compared with the Al2O3 catalyst, the Ni/Al2O3 catalyst not only has acidic sites but also active metal sites, thus showing more outstanding performance in improving the quality of bio-oil.
Santander et al. [50] used CeO2 catalysts and bifunctional Cu/Ni/ZrO2 catalysts for the catalytic conversion of sunflower seed hulls in a two-stage fixed-bed reactor and investigated their effects on the distribution and quality of pyrolysis products. The results showed that the addition of the catalyst could effectively improve the quality of the bio-oil. Compared with the absence of catalysts, the presence of the CeO2 catalyst could significantly reduce the concentration of methoxyphenols and promote the formation of alkylphenols and high molecular weight ketones, thereby leading to the reduction of the oxygen content of the biomass oil from 23.1 to 13.2 wt%, with the heating value increasing from 29.6 MJ/kg to 35.5 MJ/kg. In addition, the Cu/Ni/ZrO2 bifunctional catalyst could promote the C-C coupling reaction, demethoxylation, and methylation reactions, which further reduced the oxygen content (8.1 wt%) and increased the H/C ratio and heating value (38.7 MJ/kg) of the bio-oil.
Luo et al. [51] used deep eutectic solvent pretreatment technology and bio-enzymatic hydrolysis technology combined with catalytic conversion technology to catalyze the conversion of sugarcane bagasse to prepare bio-jet fuel. The results showed that deep eutectic solvent pretreatment could effectively destroy the biomass structure of sugarcane bagasse, which could promote bagasse enzymolysis and fermentation, and, combined with bio-enzymatic hydrolysis technology, acetone/butanol/ethanol intermediates were prepared using bagasse as a raw material. The catalytic conversion of acetone/butanol/ethanol using a single catalyst bed (2% Ni/HBET catalyst) produces a large number of ethers (mainly dibutyl ether) that severely inhibit the growth of the carbon chain, thus reducing the quality and selectivity of bio-jet fuel. In contrast, the selectivity and conversion of bio-jet fuel can be effectively improved by using a dual catalyst bed (HSAPO-34 catalyst and Ni/HBET catalyst). The microporous HSAPO-34 catalyst reduced the conversion rate of acetone/butanol/ethanol intermediates into ethers and significantly improved the selectivity of lighter olefins, and the mesoporous 2% Ni/HBET catalyst effectively promoted the synthesis of bio-jet fuel, thus improving the conversion rate of acetone/butanol/ethanol intermediates (95.3%) and selectivity of C9-C16 hydrocarbons (83.0%).

2.3. Application of Catalysts in Biomass Syngas

Bio-gas is a mixture of gases produced during biomass conversion, including carbon dioxide, carbon monoxide, hydrogen, methane, low molecular weight hydrocarbons, and some impure gases. With different raw biomass materials and conversion conditions, the compositions of bio-gas and contents of various gases are different, resulting in a wide range of potential utilization values for use as a carrier gas, in fuel heating or power generation, or as a synthetic biomass liquid fuel [52]. In the process of biomass conversion, the addition of appropriate catalysts can not only improve the yield of bio-gas, but also change the composition of bio-gas, thereby increasing the calorific value of bio-gas [53].
Zhang et al. [47] used metal oxides such as MgO, CaO, and La2O3 as catalysts for the pyrolysis and conversion of poplar, cellulose, and lignin in a fixed bed reactor. The results showed that the alkaline centers in the alkaline catalysts (CaO and MgO) could effectively promote the cleavage of C-H, C-O, or C-C bonds, thereby promoting the formation of gaseous products (CO, CO2, and CH4), and the starting time sequence of their formation was of the order CO < CO2 < CH4. It was also found that pyrolysis using cellulose as a raw material produced more CO, while pyrolysis using lignin as a raw material produced more CH4. In addition, CaO and La2O3 could react with CO2 generated during pyrolysis to form carbonates, thereby reducing the yield of CO2 in pyrolysis conversion products and affecting the catalytic activity of CaO and La2O3.
Singh et al. [54] used four kinds of agricultural crop residues, namely rice husk, straw, sugarcane bagasse, and corn cob, as mixed biomass raw materials, and pyrolyzed them in a fixed-bed reactor, while the pyrolysis steam was reformed in the presence of LaNi0.5Co0.5O3 catalyst. The research results showed that the LaNi0.5Co0.5O3 catalyst exhibited high hydrogen production activity throughout the entire reaction process. At the 5th hour of the reaction, the H2 yield was 74%, the CO2 yield was 64%, and the CO yield was 38%. However, with the increase of reaction time, at the 12th hour, the yield of H2, CO2, and CO decreased to 80%, 55%, and 35%, respectively.
Tang et al. [55] used hydrofluoric acid to modify a nickel-loaded ZSM-5 catalyst (Ni/FZ5) and investigated its performance in the catalytic conversion of toluene from corn kernels and in the steam reformation of biomass pyrolysis tar in a fixed-bed reactor. The results showed that hydrofluoric acid etching could significantly increase the porosity and specific surface area of Ni/FZ5, provide more active sites for Ni, thereby enhancing the dispersion of Ni metal and improving the catalytic activity of Ni/FZ5. Meanwhile, different Ni loadings had certain effects on the catalytic performance of Ni/FZ5. When the Ni loading was 9 wt%, Ni/FZ5 had a specific surface area of 337 m2/g and an average pore size of 2.91 nm, as well as 23 nm Ni particles and high dispersibility. Ni/FZ5 also had a toluene conversion rate of 75.6%, a hydrogen production rate of biomass volatiles of 52%, and a selectivity of 72.8%. In addition, Ni/FZ5 still had high catalytic activity for hydrogen production from BSR after 7 cycles of regeneration, with a hydrogen yield of 35.5 mmol/g and a selectivity of 67.8%.
Li et al. [56] prepared HZSM-5 catalysts (Ni-Fe/HZSM-5) with Ni loading of 8 wt% and Fe loading of 1–4 wt% and investigated the effect of Ni-Fe/HZSM-5 on the catalytic conversion of biomass tar (toluene as a model compound) to prepare syngas in a fixed-bed reactor fed with a CO2 atmosphere. The results showed that Fe loading had a significant effect on the composition of the biomass catalytic products. When only HZSM-5 was used as the catalyst, the yields of bio-gas, bio-oil and bio-char were 47.3, 25.14, and 25.43 wt%, respectively, and the average rates of H2 and CH4 release were 21.3 mL/min and 35.2 mL/min. However, when Fe loading was 4 wt%, the yields of bio-gas, bio-oil and bio-char were 63.4 wt%, 28.1 wt% and 8.5 wt%, respectively, and the H2 yield was 53.3 mL/g-toluene. The CH4 yield was maintained at 56–59 mL/g-toluene with the increase of Fe content from 1 wt% to 4 wt%. Meanwhile, the amount of the Ni-Fe/HZSM-5 catalyst also affected the distribution of catalytic conversion products. When the amount of the Ni-Fe/HZSM-5 catalyst was increased from 4% to 16%, the H2 and CH4 yields increased from 42.35 and 58.93 mL/g-toluene to 72.2 and 82.0 mL/g-toluene, respectively. In addition, the Ni-Fe catalyst, CO2, and toluene had a synergistic effect in catalytic conversion, thereby increasing the yields of H2, CO, and total syngas in the reaction products and reducing the yields of CH4 and carbon deposition. When the CO2 flow rate was 40 mL/min, the yields of H2, CO, total syngas, CH4, and carbon deposition were 224.42 mL/g-toluene, 19.07 mL/g-toluene, 411.96 mL/g-toluene, 127.12 mL/g-toluene, and 230 mg/g-toluene, respectively.
Li et al. [57] used hydrothermal and acid activation methods to activate waste aluminum float (AD) and used it as a substrate to prepare carbon–aluminum-based composite catalysts (C–ACMC) and iron-loaded carbon–aluminum composite catalysts (FeC–ACMC), and at the same time, these catalysts were applied to the catalytic conversion of poplar wood chips to investigate the effect of the catalysts on the pyrolysis of the biomass. The results showed that activated AD could effectively improve its catalytic performance, especially after acid activation, which could achieve a H2 release rate of 15.61 mL/min, a gas yield of 237.85 mL/g-biomass, a CO2 yield of 49.62 mL/g-biomass, a CO yield of 77.53 mL/g-biomass, and a CH4 yield of 5.45 mL/g-biomass. Meanwhile, the C–ACMC exhibited a synergistic effect of acid activation and carbon properties, which could effectively improve its catalytic activity. Especially, the acetic acid-activated carbon–aluminum composite catalyst (C–ACMC/AA) showed outstanding catalytic performance, which could obtain a 86.65 mL/g-biomass H2 yield and 78.34 mL/g-biomass CO yield. In addition, the FeC–ACMC also exhibited unique catalytic performance, obtaining a 43.77% gas yield, 19.79% coke deposition, and 64.24 mL/g-biomass H2 yield in the catalytic conversion.

3. Application of Catalysts in the Preparation of High-Value-Added Chemicals from Biomass and Its Derivatives

Biomass can be chemically converted not only to organic compounds, as well as hydrocarbon fuels, but also to prepare various high-value-added chemicals by chemical conversion, such as formic acid, levoglucan, 5-HMF, LA, acetyl propionic acid (LPA), 3-hydroxybutyrolactone, γ-valerolactone (GVL), furans, and furfural [58,59]. Using these platform chemical compounds as raw materials, a wide variety of chemical products can be prepared through hydrogenolysis, hydrogenation, oxidation, and hydrodeoxygenation reactions, which are widely used in pharmaceutical intermediates, fine chemicals, polymers, etc., such as 3-hydroxypropionic acid, which can be used in the production of plastic packaging, resins, fibers, etc., and as a precursor for the synthesis of acrylic acid and its derivatives [60]. LA has high energy density and can be used as a fuel additive [61]. 5-HMF can be converted into dimethylfuran, which can be used as both a solvent and a liquid fuel [62]. GVL can be converted into the precursor adipic acid of nylon, which opens up a “biological route” for nylon preparation [63]. Therefore, the development of efficient and green catalytic systems for the preparation of biomass platform compound molecules and their conversion into more versatile high-value-added chemicals and biomass-based fuels is conducive to maximizing the use of biomass resources, reducing the dependence on existing fossil resources and energy, and promoting resource recycling and sustainable development.

3.1. Application of Catalysts in the Preparation of 5-HMF and Its Derivatives from Biomass

5-HMF, as a high-value-added platform compound for biomass conversion, is widely used in the fields of materials, energy, medicine, and chemicals [64]. Its structure contains various functional groups such as furan rings, aldehydes, and alcohols, which can be used to prepare various high-value-added chemicals through oxidation, reduction, esterification, and other chemical reactions. It is a bridge between biorefineries and petroleum refineries and is also regarded as a “sleeping giant” in the field of green chemistry [65]. However, in undisturbed natural conditions, the conversion of biomass and its derivatives to prepare 5-HMF exhibits low efficiency, yield, and selectivity. Therefore, it is particularly important to design and prepare catalysts with high catalytic activity and selectivity for the green and efficient production of 5-HMF. The effect of catalysts on the preparation of HMF and its derivatives from biomass is shown in Table 2.
Cai et al. [66] used low-toxicity metal chloride salts (AlCl3, FeCl3, CoCl2, ZnCl2, and CaCl2) as catalysts to investigate their effects on the conversion of glucose to 5-HMF in a reaction solvent composed of green solvent GVL and water, combined with microwave-assisted heating technology. The results showed that the same amounts of different metal chlorides in the reaction system exhibited different acidity, which led to the lower selectivity of FeCl3 as a catalyst to catalyze the preparation of 5-HMF from glucose and the higher selectivity of AlCl3, CoCl2, ZnCl2, and CaCl2 for 5-HMF. Meanwhile, the presence of AlCl3 can effectively promote the isomerization of glucose, which has a high yield (74.04%) for the preparation of 5-HMF. In addition, the AlCl3 catalyst also showed excellent stability, and the yield of 5-HMF could still reach 68.56% when it was recycled four times.
Hirano et al. [67] hybridized different metal ions (Pt2+, Ni2+, Co2+, Fe2+) and graphene oxide (GO) to prepare a series of graphene oxide hybrid catalysts (M-rGO) and investigated their effects on the conversion of glucose to 5-HMF. The results showed that modifying GO and metal ions can prepare hybrid catalysts (M-rGO) with a huge specific surface area and form active catalytic sites with Lewis and Brønsted functional groups, thereby improving the catalytic performance of M-rGO catalysts, among which Ni-rGO showed the highest catalytic effect on 5-HMF, whose glucose conversion was 99% under the suitable reaction conditions, and the 5-HMF yield was 75%.
Wei et al. [68] used polydopamine (PDA) to modify graphene oxide and prepared a UiO-66-type metal organic framework (MOFs) catalyst (UiO-66-SO3H/PDA@GO and UiO-66-NH2/PDA@GO) with a dual functional acid–base through a one pot method. The results showed that the acidic and basic sites of the bifunctional acid–base MOFs catalysts contributed to the isomerization of glucose to fructose and the dehydration of fructose to 5-HMF, and, under the optimal reaction conditions, the UiO-66-SO3H-NH2/PDA@GO catalyst could catalyze the conversion of glucose to 5-HMF with a 55.8% yield, and the catalyst was recycled five times without any significant decrease in its catalytic efficiency.
2,5-Furan dicarboxylic acid (FDCA) and 2,5-di(hydroxymethyl)furan (BHMF), as important derivatives of the biomass platform molecule 5-HMF, are high-value-added chemicals that replace petroleum-based terephthalic acid, synthesize polyester compounds, and prepare biodiesel. They have been widely used in a variety of applications, such as pharmaceuticals, ethers, ketones, synthetic fibers, resins, and other fields [76,77]. However, a variety of by-products are generated during the synthesis of FDCA and BHMF by HMF, which reduces the yield and selectivity of the target products, yet the use of suitable catalysts is an effective way to solve this problem.
Ren et al. [71] used layered double hydroxides (LDHs) as precursors to prepare bimetallic oxide composite (CoMn-MOC) catalysts loaded with Co3O4 and Co2MnO4 and investigated their effects on the selective oxidation of 5-HMF to prepare FDCA. The results showed that the CoMn-MOC catalysts could undergo a strong localized electron exchange, which promoted the activated adsorption of C-O groups in furan and enhanced the activity of the catalyst, resulting in more than 99% conversion of 5-HMF and 98% yield of FDCA. In addition, the CoMn-MOC catalysts showed satisfactory stability and reusability, and after they were continuously recycled six times, the 5-HMF conversion was 95% and the FDCA yield remained above 94%.
Zhou et al. [72] used a cobalt/zinc lignin (Co/Zn-L) composite as a template and prepared a Co-SAs/N@C catalyst through single-atom cobalt loaded on nitrogen-doped carbon and investigated its effect on the preparation of FDCA. The results showed that dispersed Co atoms on nitrogen-doped carbon atoms had excellent reactivity and selectivity for the oxidation of alcohols and aldehydes to carboxylic acids, and Co-SAs/N@C catalyst for the catalytic conversion of furfural and CO2 to prepare FDCA had a 76% yield and 82.6% selectivity. In addition, Co-SAs/N@C could catalyze the synthesis of FDCA from 5-HMF, and the conversion of 5-HMF reached 99.4% after 3 h of reaction, and when the reaction was carried out for 8 h, the yield of FDCA was close to the theoretical value (99.5%).
Pan et al. [73] prepared a Ni(OH)2/SiO2 composite catalyst by loading Ni(OH)2 onto SiO2 carriers using the vapor-induced internal hydrolysis method and investigated its effect on the preparation of 2,5-bis(hydroxymethyl)furan (BHHMF) by the hydrogenation of 5-HMF. The results showed that the Ni(OH)2/SiO2 composite catalyst did not need other additional pretreatment to exhibit excellent catalytic performance for the hydrogenation of 5-HMF to BHMF, and the catalytic activity of the Ni(OH)2/SiO2 composite catalyst could be effectively enhanced by adjusting the relative ratio between the active component and the carrier. When the Ni content was about 5 wt%, the 5-HMF conversion rate was 87.6% and the BHMF selectivity was 98.9%. In addition, the Ni(OH)2/SiO2 composite catalyst showed excellent stability, and the 5-HMF conversion rate was still more than 90% when it was recycled four times.

3.2. Application of Catalysts in the Preparation of LA and Its Derivatives from Biomass

LA is one of the high-value-added sustainable platform chemicals that can be obtained through the catalytic conversion of biomass which contains both carbonyl and carboxylic acid groups. It can undergo a variety of chemical reactions, such as hydrolysis, oxidation, hydrogenation, halogenation, and condensation, and has been used in a wide range of applications such as biofuels, pharmaceuticals, lubricants, and fine chemicals [78]. The hydrolysis of furfuryl alcohol and direct conversion of biomass and its derivatives are the main methods for the preparation of LA, but in the absence of catalysts, these two methods resulted in low yields of LA, accompanied by a large number of by-products, which seriously affected the high-value application of LA [79]. Therefore, the use of suitable catalysts is an indispensable and effective method and pathway to improve the yield and selectivity of LA [80]. The effects of catalysts on the preparation of LA and its derivatives from biomass are shown in Table 3.
Zhang et al. [81] prepared a series of bifunctional catalysts (5Zr-HY, 10Zr-HY, and 15Zr-HY) with an octahedral morphology and good stability by loading different contents of metal Zr on HY zeolite carriers and investigated their effects on the conversion of glucose to LA. The results showed that Zr was loaded onto the HY zeolite skeleton in the form of a tetrahedral coordination and formed bifunctional catalysts with Lewis and Brønsted acid sites, which effectively improved its catalytic performance. Under suitable reaction conditions, the conversion rate of glucose and the yield of LA could reach 100% and 43.0%, respectively. Meanwhile, the catalyst still showed good stability and reusability after repeated recycling four times.
Xu et al. [82] prepared a carbon foam catalyst (HAlW/CF) loaded with H5AlW12O40 by a hydrothermal impregnation method and investigated its effect on the direct catalytic conversion of cellulose to synthesize LA. The results showed that the prepared HAlW/CF catalyst possessed the bis-acidic sites of Brønsted and Lewis and exhibited high total acidity. In the process of the catalytic conversion of cellulose to produce LA, carbon foam (CF) and aluminotungstic acid (HAlW) showed good synergistic effects, which improved the catalytic activity and selectivity of the HALW/CF catalyst. The conversion of cellulose could reach 89.4%, and the yield of LA was 60.9%. In addition, the HAlW/CF catalysts were easy to separate, recycle, and reuse. After five consecutive cycles of recycling, the conversion rate of cellulose and the yield of LA were 80.6% and 50.6%, respectively.
Azlan et al. [83] used phosphotungstic acid (HPW) and niobium oxide to prepare a bifunctional catalyst (HPWNb2O5) with simultaneous Brønsted and Lewis acids and investigated its effect on the preparation of LA from peel fibers in oil palm. The results showed that the simultaneous presence of Brønsted acid and Lewis acid provides synergistic effects for the conversion of biomass to LA, and compared with individually functionalized catalysts (HPW, H2SO4, InCl3, Al(OTf)3, Bi(OTf)3, and Nb2O5), the HPWNb2O5 bifunctional catalyst has a higher yield (16.14%) for the synthesis of LA from biomass. After doping in lignin-derived carbon cryogels, it showed strong stability and reusability, and when recycled three times, its yield for LA was still 13.1–14.1%.
As an important derivative of levulinic acid (LA), GVL can be widely used as an environmentally friendly solvent for biorefineries, an additive for energy-intensive fuels, and an intermediate for high-added chemicals due to its easy conversion to other chemicals [89]. In the process of converting LA to GVL, a suitable and efficient catalyst is particularly important to improve the yield and selectivity of GVL [90,91].
Xu et al. [86] prepared carbon-loaded Ni and Co bimetallic catalysts (Ni1Co1/C) by pyrolysis using NiCo-BTC metal-organic skeleton (MOF) materials as precursors and used them for the application of biomass-derived LA selective hydrogenation for the preparation of GVL. The results showed that the porous Ni1Co1/C catalyst carbon carrier has a large specific surface area and provides abundant active sites for Ni and Co bimetals. At the same time, the dilution and separation of metal Ni active sites by Co results in the enhanced adsorption of NiCo metal sites, which leads to higher catalytic activity and stability for Ni1Co1/C bimetallic catalysts than Ni/C and Co/C monometallic catalysts, and the Ni1Co1/C catalysts can make LA almost completely converted, with a GVL yield of 95.2%. The yield loss of GVL was no more than 3% after continuous recycling seven times.
Meng et al. [74] loaded cucurbit[5]uirl (CB[5]) onto semiconducting tungsten trioxide (Ov-WO3) with oxygen vacancies, thereby preparing a novel host–guest photocatalyst (Ov-WO3@CB[5]), which was used to reduce the modified biomass derivative ethyl acetylpropionate. The results showed that the introduction of oxygen vacancies and CB[5] can significantly improve the activity of the Ov-WO3@CB[5] catalyst, under the appropriate conditions, and the catalytic conversion of ethyl acetylpropionate to GVL was achieved in a 99% yield with the catalyst Ov-WO3@CB[5], while the yield of the WO3 catalyst was only 41%. Meanwhile, the Ov-WO3@CB[5] catalyst also showed excellent stability, and the yield of GVL was still as high as 91% after five consecutive cycles.
Yang et al. [87] prepared SiO2@RuO2@RF catalysts with a yolk–shell structure (Ru@me-HNC) by depositing RuO2 nanoparticles on NH2-modified SiO2 nanorods, followed by resorcinol-formaldehyde (RF) resin as a shell. Ru/me HNC and Ru/activated carbon (Ru/AC) catalysts with external Ru loading were synthesized by the wet impregnation reduction method and their effects on the upgrading of LA to GVL were investigated. The results showed that the prepared SiO2@RuO2@RF catalysts possessed an ultra-high specific surface area (4016 m2/g), which enhanced the Ru0-H interactions and electron transfer between C-O groups and thus exhibited strong catalytic activity. Under appropriate reaction conditions, the selectivity of GVL was 99.9%, and the yield was 98.2%.
Zhao et al. [88] prepared a ZrP/SBA-15 catalyst by impregnating the synthesized ZrO2/SBA-15 catalyst with a phosphoric acid solution and used it to catalyze the catalytic conversion of acetylpropionic acid to GVL. The results showed that the modification of ZrO2 metal oxides using phosphoric acid could selectively form the activated zirconium phosphates, thereby changing the acidic sites of Brønsted (B) and Lewis (L), which showed excellent catalytic activity and stability, with the conversion of LA exceeding 98% and the selectivity of GVL over 61%. The catalytic activity was still high even after three cycles of reuse and recycling.

4. Prospects and Challenges of Catalysts in Biomass Conversion

The chemical conversion of biomass is an effective way to achieve efficient and sustainable utilization of biomass resources and energy. It is an important bridge to communicate the high-value-added conversion and substitution of biomass resources for fossil resources, which plays an indispensable role in coping with shortages of fossil resources and energy, increases in environmental pollution, and climate change in many aspects. As an important influencing factor in the efficient conversion of biomass, catalysts have made positive contributions in reducing the activation energy of chemical reactions, improving the yield and selectivity of reaction products and optimizing reaction conditions. However, insufficient research on the mechanisms of catalyst-catalyzed biomass conversion, further enhancement of catalyst-specific selection of target products, efficient conversion efficiency, deactivation, regeneration of catalysts, and separation and purification of catalytic products have all become challenges for efficient and green biomass conversion. Therefore, in future biomass conversion processes, we need to adopt more scientific and advanced methods to conduct more pragmatic and in-depth research on the application of catalysts in biomass conversion. This includes the following:
(1)
Thorough study of the reaction mechanisms and kinetics of catalysts in biomass conversion. People should adopt advanced scientific research methods and technologies to further systematically deepen and elucidate the key theoretical research of catalysts in each step of biomass conversion processes, especially in the study of efficient catalytic mechanisms for specific biomass conversion products and high-value-added chemical components, in order to better guide practical applications with theory.
(2)
Focusing on the catalytic efficiency, deactivation, regeneration, and recycling of catalysts in the biomass conversion process. Adopting new technologies and concepts to develop high-performance green catalysts with high selectivity, good stability, long service life, low cost, and environmental friendliness.
(3)
Optimizing and constructing the reaction conditions and production processes for catalysts in biomass conversion. Adopting scientific and systematic methods to construct safe, stable, economical, efficient, and green biomass conversion processes, reasonably optimizing and adjusting different reaction conditions, and achieving their synergistic effect on the catalytic conversion of biomass.
(4)
Exploring the application of catalysts in biomass conversion by combining structure-reactivity relationships on the basis of molecular simulation and machine learning. Through new technologies such as molecular simulation and machine learning, we can scientifically integrate experimental and computational data, where, combined with structure reactivity relationships, efficient catalysts are directly screened and specifically designed and their technical and economic performance is reasonably evaluated.

5. Conclusions

With the continuous consumption of fossil resources and the continuous deterioration of the ecological environment, it is particularly urgent to find renewable green resources that can replace fossil resources. Biomass, as a renewable energy source that directly converts light energy into chemical energy and exists in the form of carbon containing substances, can catalyze its conversion to produce biomass energy and high-value-added chemicals, thus demonstrating enormous potential in replacing fossil resources and sustainable development. Therefore, this article focuses on the resource utilization of lignocellulosic biomass in the fields of chemical conversion and high-value-added chemical preparation and provides a detailed overview of the application of catalysts in biomass energy, represented by bio-char, bio-oil, and bio-gas, as well as high-value-added chemicals and their derivatives, represented by 5-HMF and LA. At the same time, the difficulties and challenges encountered by catalysts in biomass conversion are also analyzed. In the chemical conversion of biomass, catalysts can effectively increase the yield of reaction products, improve the quality of reaction products, and enhance the selectivity of target products. Metal catalysts and modified catalysts are commonly used catalysts to effectively promote biomass catalytic conversion, which has a significant impact on the composition and distribution of bio-oil and bio-gas products. Composite catalysts and bifunctional catalysts can adjust the acidic sites of Lewis and Brønsted in catalysts according to different reaction requirements, thereby improving the targeted reaction efficiency, yield, and selectivity of specific chemical products, which has become a widely studied hotspot. However, catalysts were also found to have challenges and problems in the biomass conversion process, such as vague catalytic mechanisms, being prone to losing catalytic activity, complicated regeneration processes, and difficulty in the separation and purification of catalytic products. Therefore, in future biomass catalytic conversion processes, we should conduct more practical and scientific research from multiple aspects, such as in-depth research on the reaction mechanism and kinetics of catalysts in biomass conversion, focusing on the catalytic efficiency, deactivation, regeneration, and recovery of catalysts in biomass conversion, as well as optimizing and constructing the reaction conditions and production processes of catalysts in biomass conversion and exploring the application of catalysts in biomass conversion through structure-reactivity relationships, molecular simulation and machine learning, so as to make active contributions to the design and development of catalysts with high efficiency, greenness, and sustainability.

Author Contributions

Conceptualization, J.C., X.Z. and W.L.; methodology, J.C. and L.W.; software, J.C. and F.L.; validation, J.C. and Y.L.; formal analysis, J.C., J.W. and N.L.; investigation, L.W.; resources, J.C. and X.Z.; data curation, L.W.; writing—original draft preparation, J.C., X.Z. and W.L.; writing—review and editing, L.W. and J.C.; visualization, J.W., N.L. and Y.L.; supervision, X.Z. and W.L.; project administration, F.L., X.Z. and W.L.; funding acquisition, J.C. and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the Natural Science Foundation of the Xinjiang Uygur Autonomous Region (2022D01B02).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Theerthagiri, J.; Karuppasamy, K.; Park, J.; Rahamathulla, N.; Kumari, M.L.A.; Souza, M.K.R.; Cardoso, E.S.F.; Murthy, A.P.; Maia, G.; Kim, H.S.; et al. Electrochemical conversion of biomass-derived aldehydes into fine chemicals and hydrogen: A review. Environ. Chem. Lett. 2023, 21, 1555–1583. [Google Scholar] [CrossRef]
  2. Babu, S.; Jojo, L.; James, A.; Melethil, K.; Thomas, B. Metal-based catalysis for biomass and renewables valorization-current status. Tetrahedron Green Chem 2023, 2, 100018. [Google Scholar] [CrossRef]
  3. Guo, L.Y.; Li, J.F.; Lu, Z.W.; Zhang, J.; He, C.T. Biomass-Derived Carbon-Based Multicomponent Integration Catalysts for Electrochemical Water Splitting. ChemSusChem 2023, 16, e202300214. [Google Scholar] [CrossRef] [PubMed]
  4. Cai, J.; Lin, N.; Li, Y.; Xue, J.; Li, F.; Wei, L.; Yu, M.; Zha, X.; Li, W. Research on the application of catalytic materials in biomass pyrolysis. J. Anal. Appl. Pyrolysis. 2024, 177, 106321. [Google Scholar] [CrossRef]
  5. Samoylova, Y.V.; Sorokina, K.N.; Parmon, V.N. Use of Microalgae Biomass to Synthesize Marketable Products: 2. Modern Approaches to Integrated Biorefinery of Microalgae Biomass. Catal. Industry 2024, 16, 69–76. [Google Scholar] [CrossRef]
  6. Wu, G.; Shen, C.; Liu, S. Research progress on the preparation and application of biomass derived methyl levulinate. Green Chem. 2021, 23, 9254–9282. [Google Scholar] [CrossRef]
  7. Delmas, G.H.; Banoub, J.H.; Delmas, M. Lignocellulosic biomass refining: A review promoting a method to produce sustainable hydrogen, fuels, and products. Waste Biomass Valorization 2022, 13, 2477–2491. [Google Scholar] [CrossRef]
  8. Dwivedi, A.D.; Priya, B.; Chinthala, R. Valorization of biomass-derived furans over molecular catalysts. Tetrahedron Green Chem 2023, 1, 100008. [Google Scholar] [CrossRef]
  9. Yogalakshmi, K.N.; Sivashanmugam, P.; Kavitha, S.; Kannah, Y.; Varjani, S.; AdishKumar, S. Lignocellulosic biomass-based pyrolysis: A comprehensivereview. Chemosphere 2022, 286, 131824. [Google Scholar]
  10. Wang, S.; Li, Z.; Bai, X.; Yi, W.; Fu, P. Catalytic pyrolysis of lignin in a cascade dual-catalyst system of modified red mud and HZSM-5 for aromatic hydrocarbonproduction. Bioresour. Technol. 2019, 278, 66–72. [Google Scholar] [CrossRef] [PubMed]
  11. Nanda, S.; A Kozinski, J.; K Dalai, A. Lignocellulosic biomass: A review of conversion technologies and fuel products. Curr. Biochem. Eng. 2016, 3, 24–36. [Google Scholar] [CrossRef]
  12. Fan, K.; Mao, W.; Qu, H. Study on government subsidy in a two-level supply chain of direct-fired biomass power generation based on contract coordination. J. Ind. Manag. Optim. 2023, 19, 2436–2451. [Google Scholar] [CrossRef]
  13. Meena, M.K.; Anand, S.; Ojha, D.K. Interdependency of pyrolysis and combustion: A case study for lignocellulosic biomass. J. Therm. Anal. Calorim. 2023, 148, 5509–5519. [Google Scholar] [CrossRef]
  14. Pati, S.; De, S.; Chowdhury, R. Exploring the hybrid route of bio-ethanol production via biomass co-gasification and syngas fermentation from wheat straw and sugarcane bagasse: Model development and multi-objective optimization. J. Clean. Product. 2023, 395, 136441. [Google Scholar] [CrossRef]
  15. Manikandan, S.; Vickram, S.; Sirohi, R. Critical Review of Biochemical Pathways to Transformation of Waste and Biomass into Bioenergy. Bioresour. Technol. 2023, 372, 128679. [Google Scholar] [CrossRef] [PubMed]
  16. Deng, W.; Feng, Y.; Fu, J. Catalytic conversion of lignocellulosic biomass into chemicals and fuels. Green Energy Environ. 2023, 8, 10–114. [Google Scholar]
  17. Chen, J.Y.; Xiao, Y.; Guo, F.S.; Li, K.M.; Huang, Y.B.; Lu, Q. Single-Atom Metal Catalysts for Catalytic Chemical Conversion of Biomass to Chemicals and Fuels. ACS Catal. 2024, 14, 5198–5226. [Google Scholar] [CrossRef]
  18. Deng, W.; Wang, Y. Research perspectives for catalytic valorization of biomass. J. Energy Chem. 2023, 78, 102–104. [Google Scholar] [CrossRef]
  19. Nzediegwu, E.; Dumont, M.J. Chemo-catalytic transformation of cellulose and cellulosic-derived waste materials into platform chemicals. Waste Biomass Valorization 2021, 12, 2825–2851. [Google Scholar]
  20. Peng, L.; Wang, M.; Li, H. Tert-Butanol intervention enables chemoselective conversion of xylose to furfuryl alcohol over heteropolyacids. Green Chem. 2020, 22, 5656–5665. [Google Scholar] [CrossRef]
  21. Rahman, M.M.; Liu, R.; Cai, J. Catalytic fast pyrolysis of biomass over zeolites for high quality bio-oil—A review. Fuel Process. Technol. 2018, 180, 32–46. [Google Scholar] [CrossRef]
  22. Shetty, A.; Molahalli, V.; Sharma, A. Biomass-derived carbon materials in heterogeneous catalysis: A step towards sustainable future. Catalysts 2022, 13, 20. [Google Scholar] [CrossRef]
  23. Yan, Q.; Wu, X.; Jiang, H. Transition metals-catalyzed amination of biomass feedstocks for sustainable construction of N-heterocycles. Coord. Chem. Rev. 2024, 502, 215622. [Google Scholar] [CrossRef]
  24. Lodhi, A.; Maheria, K.C. Zeolite-catalysed esterification of biomass-derived acids into high-value esters products: Towards sustainable chemistry. Catal. Commun. 2024, 187, 106883. [Google Scholar] [CrossRef]
  25. Auta, M.; Ern, L.M.; Hameed, B.H. Fixed-bed catalytic and non-catalytic empty fruit bunch biomass pyrolysis. J. Anal. Appl. Pyrolysis 2014, 107, 67–72. [Google Scholar] [CrossRef]
  26. Hwang, H.; Oh, S.; Choi, I.G.; Choi, J.W. Catalytic effects of magnesium on the characteristics of fast pyrolysis products–Bio-oil, bio-char, and non-condensed pyrolytic gas fractions. J. Anal. Appl. Pyrolysis 2015, 113, 27–34. [Google Scholar] [CrossRef]
  27. Mohamed, B.A.; Kim, C.S.; Ellis, N. Microwave-assisted catalytic pyrolysis of switchgrass for improving bio-oil and biochar properties. Bioresour. Technol. 2016, 201, 121–132. [Google Scholar] [CrossRef] [PubMed]
  28. Chong, Y.Y.; Thangalazhy-Gopakumar, S.; Ng, H.K. Effect of oxide catalysts on the properties of bio-oil from in-situ catalytic pyrolysis of palm empty fruit bunch fiber. J. Environ. Manag. 2019, 247, 38–45. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, X.; Chen, Y.; Yang, H. Fast pyrolysis of cotton stalk biomass using calcium oxide. Bioresour. Technol. 2017, 233, 15–20. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, C.T.; Zhang, L.; Li, Q. Catalytic pyrolysis of poplar wood over transition metal oxides: Correlation of catalytic behaviors with physiochemical properties of the oxides. Biomass Bioenergy 2019, 124, 125–141. [Google Scholar] [CrossRef]
  31. Dong, Q.; Li, H.; Niu, M. Microwave pyrolysis of moso bamboo for syngas production and bio-oil upgrading over bamboo-based biochar catalyst. Bioresour. Technol. 2018, 266, 284–290. [Google Scholar] [CrossRef] [PubMed]
  32. Guo, F.; Dong, Y.; Dong, K. Role of biochar-based catalysts in microwave-induced biomass pyrolysis: Structural properties and modification with Fe-series metals. Fuel 2023, 341, 127769. [Google Scholar] [CrossRef]
  33. Li, X.; Fu, H.; Shao, S. Synthesis of hierarchical HZSM-5 utilizing natural cellulose as templates for promoted production of aromatic hydrocarbons in the catalytic pyrolysis of biomass. Fuel Process. Technol. 2023, 248, 107815. [Google Scholar] [CrossRef]
  34. Naqvi, S.R.; Naqvi, M.; Inayat, A. Impact of layered and delaminated zeolites on catalytic fast pyrolysis of microalgae using fixed-bed reactor and Py-GC/MS. J. Anal. Appl. Pyrolysis 2021, 155, 105025. [Google Scholar] [CrossRef]
  35. Gai, D.; Shi, J.; Cui, X. Catalytic performance and mechanism of A-site vacancy deficient perovskite catalyst over tar cracking during biomass pyrolysis. J. Clean. Prod. 2023, 405, 136876. [Google Scholar] [CrossRef]
  36. Hubble, A.H.; Childs, B.A.; Pecchi, M. Role of in situ (in contact with biomass) and ex situ (in contact with pyrolysis vapors) transition metal catalysts on pyrolysis of cherry pits. Fuel 2023, 352, 129062. [Google Scholar] [CrossRef]
  37. Pagano, M.; Hernando, H.; Cueto, J. Assessing the changes in thermal and catalytic pyrolysis of lignocellulose when shifting from batch to continuous biomass feeding modes. Catal. Today 2024, 426, 114399. [Google Scholar] [CrossRef]
  38. Qiu, B.; Tao, X.; Wang, J. Research progress in the preparation of high-quality liquid fuels and chemicals by catalytic pyrolysis of biomass: A review. Energy Convers. Manag. 2022, 261, 115647. [Google Scholar] [CrossRef]
  39. Liu, W.; Wu, X. Ligin-derived graphene-like ultrathin carbon materials for ORR catalysis. J. Phys. Conf. Ser. 2024, 2749, 012003. [Google Scholar]
  40. Kong, L.; Wang, J.; Dong, K. Catalytic pyrolysis characteristics of polystyrene by biomass char-supported nanocatalysts. J. Anal. Appl. Pyrolysis 2024, 179, 106511. [Google Scholar] [CrossRef]
  41. Xu, H.; Jiang, W.; Wang, D. Controlling the growth of ZnNiMoSx on biomass-derived porous carbon materials for deep hydrodesulfurization. Appl. Catal. A General 2023, 660, 119188. [Google Scholar] [CrossRef]
  42. Lu, D.; Yang, Q.; Chen, Z. High-efficiency phenol removal by novel biomass-based alginate composite hydrogel. Chem. Phys. Lett. 2023, 826, 140676. [Google Scholar] [CrossRef]
  43. Hu, X.; Gholizadeh, M. Biomass pyrolysis: A review of the process development and challenges from initial researches up to the commercialisation stage. J. Energy Chem. 2019, 39, 109–143. [Google Scholar] [CrossRef]
  44. Hassan, N.S.; Jalil, A.A.; Hitam, C.N.C. Biofuels and renewable chemicals production by catalytic pyrolysis of cellulose: A review. Environ. Chem. Lett. 2020, 18, 1625–1648. [Google Scholar] [CrossRef]
  45. Hoang, A.T. 2-Methylfuran (MF) as a potential biofuel: A thorough review on the production pathway from biomass, combustion progress, and application in engines. Renew. Sustain. Energy Rev. 2021, 148, 111265. [Google Scholar] [CrossRef]
  46. Liu, R.; Rahman, M.M.; Sarker, M. A review on the catalytic pyrolysis of biomass for the bio-oil production with ZSM-5: Focus on structure. Fuel Process. Technol. 2020, 199, 106301. [Google Scholar]
  47. Zhang, C.; Hu, X.; Guo, H. Pyrolysis of poplar, cellulose and lignin: Effects of acidity and alkalinity of the metal oxide catalysts. J. Anal. Appl. Pyrolysis 2018, 134, 590–605. [Google Scholar] [CrossRef]
  48. Li, X.; Sun, J.; Shao, S. Preparation of composite HZSM-5 catalyst by green template and catalytic the pyrolysis of biomass to produce aromatics. Renew. Energy 2023, 206, 506–513. [Google Scholar] [CrossRef]
  49. Gupta, S.; Mondal, P. Catalytic pyrolysis of pine needles with nickel doped gamma-alumina: Reaction kinetics, mechanism, thermodynamics and products analysis. J. Clean. Prod. 2021, 286, 124930. [Google Scholar] [CrossRef]
  50. Santander, J.A.; Diez, A.; Gutierrez, V. Partial deoxygenation of pyrolysis intermediates toward fossil fuel miscible bioliquids. Environ. Prog. Sustain. Energy 2023, 42, e13945. [Google Scholar] [CrossRef]
  51. Luo, Y.; Zhang, R.; He, Y. Preparation of bio-jet fuels by a controllable integration process: Coupling of biomass fermentation and olefin polymerization. Bioresour. Technol. 2023, 382, 129175. [Google Scholar] [CrossRef] [PubMed]
  52. Kan, T.; Strezov, V.; Evans, T. Catalytic pyrolysis of lignocellulosic biomass: A review of variations in process factors and system structure. Renew. Sustain. Energy Rev. 2020, 134, 110305. [Google Scholar] [CrossRef]
  53. Xu, Z.; Gao, N.; Ma, Y. Biomass volatiles reforming by integrated pyrolysis and plasma-catalysis system for H2 production: Understanding roles of temperature and catalyst. Energy Convers. Manag. 2023, 288, 117159. [Google Scholar] [CrossRef]
  54. Singh, P.P.; Jaswal, A.; Singh, A. From Waste to Clean Energy: An Integrated Pyrolysis and Catalytic Steam Reforming Process for Green Hydrogen Production from Agricultural Crop Residues. ACS Sustain. Chem. Eng. 2024, 12, 2058–2069. [Google Scholar] [CrossRef]
  55. Tang, W.; Cao, J.P.; Yang, F.L. Highly active and stable HF acid modified HZSM-5 supported Ni catalysts for steam reforming of toluene and biomass pyrolysis tar. Energy Convers. Manag. 2020, 212, 112799. [Google Scholar] [CrossRef]
  56. Li, X.; Liu, P.; Huang, S. Study on the mechanism of syngas production from catalytic pyrolysis of biomass tar by Ni–Fe catalyst in CO2 atmosphere. Fuel 2023, 335, 126705. [Google Scholar] [CrossRef]
  57. Li, X.; Chen, Z.; Liu, P. Feasibility assessment of recycling waste aluminum dross as a basic catalyst for biomass pyrolysis to produce hydrogen-rich gas. Int. J. Hydrogen Energy 2023, 48, 36361–36376. [Google Scholar]
  58. Tong, X.; Zheng, J.; Xue, S. High-Entropy Materials as the Catalysts for Valorization of Biomass and Biomass-Derived Platform Compounds. ACS Sustain. Chem. Eng. 2023, 11, 10203–10218. [Google Scholar] [CrossRef]
  59. Agarwal, B.; Kailasam, K.; Sangwan, R.S. Traversing the history of solid catalysts for heterogeneous synthesis of 5-hydroxymethylfurfural from carbohydrate sugars: A review. Renew. Sustain. Energy Rev. 2018, 82, 2408–2425. [Google Scholar] [CrossRef]
  60. Liu, B.; Xiang, S.; Zhao, G. Efficient production of 3-hydroxypropionate from fatty acids feedstock in Escherichia coli. Metab. Eng. 2019, 51, 121–130. [Google Scholar] [CrossRef] [PubMed]
  61. Jiang, Y.; Yang, L.; Bohn, C.M. Speciation and kinetic study of iron promoted sugar conversion to 5-hydroxymethylfurfural (HMF) and levulinic acid (LA). Org. Chem. Front. 2015, 2, 1388–1396. [Google Scholar] [CrossRef]
  62. Román-Leshkov, Y.; Barrett, C.J.; Liu, Z.Y. Production of dimethylfuran for liquid fuels from biomass-derived carbohydrates. Nature 2007, 447, 982–985. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, Y.; Yuan, X.; Liu, J. Recent Advances in Zeolites-Catalyzed Biomass Conversion to Hydroxymethylfurfural: The Role of Porosity and Acidity. ChemPlusChem 2024, 89, e202300399. [Google Scholar] [CrossRef] [PubMed]
  64. Hu, L.; Lin, L.; Wu, Z. Recent advances in catalytic transformation of biomass-derived 5-hydroxymethylfurfural into the innovative fuels and chemicals. Renew. Sustain. Energy Rev. 2017, 74, 230–257. [Google Scholar] [CrossRef]
  65. Xu, C.; Paone, E.; Rodríguez-Padrón, D. Recent catalytic routes for the preparation and the upgrading of biomass derived furfural and 5-hydroxymethylfurfural. Chem. Soc. Rev. 2020, 49, 4273–4306. [Google Scholar] [CrossRef] [PubMed]
  66. Cai, J.; Xu, C.; Shi, W. Selective HMF synthesis from glucose via microwave-assisted metal chloride catalysis. Biomass Bioenergy 2024, 181, 107060. [Google Scholar] [CrossRef]
  67. Hirano, Y.; Beltramini, J.N.; Mori, A. Microwave-assisted catalytic conversion of glucose to 5-hydroxymethylfurfural using “three dimensional” graphene oxide hybrid catalysts. RSC Adv. 2020, 10, 11727–11736. [Google Scholar] [CrossRef] [PubMed]
  68. Wei, Y.; Zhang, Y.; Li, B. Fabrication of graphene oxide supported acid–based bifunctional metal–organic frameworks as efficient catalyst for glucose to 5-hydroxymethylfurfural conversion. Energy Technol. 2020, 8, 1901111. [Google Scholar] [CrossRef]
  69. García-Sancho, C.; Fúnez-Núñez, I.; Moreno-Tost, R. Beneficial effects of calcium chloride on glucose dehydration to 5-hydroxymethylfurfural in the presence of alumina as catalyst. Appl. Catal. B Environ. 2017, 206, 617–625. [Google Scholar] [CrossRef]
  70. Hu, L.; Sun, Y.; Lin, L. Efficient conversion of glucose into 5-hydroxymethylfurfural by chromium (III) chloride in inexpensive ionic liquid. Ind. Eng. Chem. Res. 2012, 51, 1099–1104. [Google Scholar] [CrossRef]
  71. Ren, Z.; Yang, Y.; Wang, S. Heterogeneous interface catalysts with electron local exchange toward highly selective oxidation of biomass platform compounds. ACS Catal. 2023, 13, 5665–5677. [Google Scholar] [CrossRef]
  72. Zhou, H.; Xu, H.; Wang, X. Convergent production of 2, 5-furandicarboxylic acid from biomass and CO2. Green Chem. 2019, 21, 2923–2927. [Google Scholar] [CrossRef]
  73. Pan, W.; Chen, H.; Chen, A. Selective transformation of biomass-derived 5-hydroxymethylfurfural to the building blocks 2, 5-bis (hydroxymethyl) furan on α-Ni (OH)2/SiO2 catalyst without prereduction. React. Kinet. Mech. Catal 2024, 137, 1635–1649. [Google Scholar] [CrossRef]
  74. Meng, Y.; Jian, Y.; Li, J. Surface-active site engineering: Synergy of photo-and supermolecular catalysis in hydrogen transfer enables biomass upgrading and H2 evolution. Chem. Eng. J. 2023, 452, 139477. [Google Scholar] [CrossRef]
  75. Ratrey, G.; Solanki, B.S.; Kamble, S.P. Highly Efficient Chemoselective Hydrogenation of 5-HMF to BHMF over Reusable Bimetallic Pd-Ir/C Catalyst. ChemistrySelect 2022, 7, e202200456. [Google Scholar] [CrossRef]
  76. Wang, L.; Zuo, J.; Zhang, Q. Catalytic transfer hydrogenation of biomass-derived 5-hydroxymethylfurfural into 2,5-dihydroxymethylfuran over Co/UiO-66-NH2. Catal. Lett. 2022, 152, 361–371. [Google Scholar] [CrossRef]
  77. Kar, S.; Zhou, Q.Q.; Ben-David, Y. Catalytic furfural/5-hydroxymethyl furfural oxidation to furoic acid/furan-2, 5-dicarboxylic acid with H2 production using alkaline water as the formal oxidant. J. Am. Chem. Soc. 2022, 144, 1288–1295. [Google Scholar] [CrossRef] [PubMed]
  78. Pachapur, V.L.; Castillo, M.V.; Saini, R. Integrated biorefinery approach for utilization of wood waste into levulinic acid and 2-Phenylethanol production under mild treatment conditions. J. Biotechnol. 2024, 389, 78–85. [Google Scholar] [CrossRef] [PubMed]
  79. Gundekari, S.; Kalusulingam, R.; Varkolu, M. Preparation of 4, 4-bis (5-methylthiophen-2-yl) pentanoic acid from biomass-based functional molecules using solid acid catalysts. Catal. Commun. 2024, 187, 106906. [Google Scholar] [CrossRef]
  80. Zhang, R.; Zhang, W.; Jiang, J. Catalytic valorization of biomass carbohydrates into levulinic acid/ester by using bifunctional catalysts. Renew. Energy 2024, 221, 119851. [Google Scholar] [CrossRef]
  81. Zhang, R.; He, S.; Wang, F. A porous B/L acid zeolite-based catalyst regulates and controls the degradation of biomass carbohydrates to produce levulinic acid/esters. Chem. Phys. 2024, 584, 112344. [Google Scholar] [CrossRef]
  82. Xu, X.; Liang, B.; Zhu, Y. Direct and efficient conversion of cellulose to levulinic acid catalyzed by carbon foam-supported heteropolyacid with Brønsted–Lewis dual-acidic sites. Bioresource Technology: Biomass, Bioenergy, Biowastes, Conversion Technologies, Biotransformations. Prod. Technol. 2023, 387, 129600. [Google Scholar]
  83. Azlan, N.S.M.; Yap, C.L.; Tiong, Y.W. The interplay of Bronsted-Lowry and Lewis Acid Sites in Bifunctional Catalyst for the Biomass Conversion to Levulinic Acid. J. Biomim. Biomater. Biomed. Eng. 2023, 61, 71–76. [Google Scholar] [CrossRef]
  84. Sobuś, N.; Czekaj, I. Catalytic transformation of biomass-derived glucose by one-pot method into levulinic acid over Na-BEA zeolite. Processes 2022, 10, 223. [Google Scholar] [CrossRef]
  85. Taghavi, S.; Ghedini, E.; Menegazzo, F. CuZSM-5@ HMS composite as an efficient micro-mesoporous catalyst for conversion of sugars into levulinic acid. Catal. Today 2022, 390, 146–161. [Google Scholar] [CrossRef]
  86. Xu, H.; Hu, D.; Lin, L. MOF-derived bimetallic NiCo nanoalloys for the hydrogenation of biomass-derived levulinic acid to γ-valerolactone. AIChE J. 2023, 69, e17973. [Google Scholar] [CrossRef]
  87. Yang, J.; Shi, R.; Xu, X. Yolk-shell electron-rich Ru@ hollow pyridinic-N-doped carbon nanospheres with tunable shell thickness and ultrahigh surface area for biomass-derived levulinic acid hydrogenation under mild conditions. Appl. Catal. B Environ. Energy 2024, 355, 124193. [Google Scholar] [CrossRef]
  88. Zhao, R.; Bae, J.W. Catalytic conversion of biomass-derived levulinic acid to γ-valerolactone over phosphate modified ZrO2/SBA-15. Catal. Today 2024, 435, 114718. [Google Scholar] [CrossRef]
  89. Martínez, F.K.G.; Martínez, F.A.; Segobia, D.J. Valeric Biofuels from Biomass-Derived γ-Valerolactone: A Critical Overview of Production Processes. ChemPlusChem 2023, 88, e202300381. [Google Scholar] [CrossRef] [PubMed]
  90. Zhao, Z.; Yang, C.; Sun, P. Synergistic catalysis for promoting ring-opening hydrogenation of biomass-derived cyclic oxygenates. ACS Catal. 2023, 13, 5170–5193. [Google Scholar] [CrossRef]
  91. Xiong, Y.; Jia, Y.; Yang, R. Foam-Like Multi-Stage Porous Al-KIT-6-4T (WI) Catalyst for Efficient Catalyzing Biomass-Derived GVL to Butene. Catal. Lett. 2024, 154, 3025–3034. [Google Scholar] [CrossRef]
Figure 1. The main components of biomass.
Figure 1. The main components of biomass.
Catalysts 14 00499 g001
Figure 2. The utilization pathways of biomass.
Figure 2. The utilization pathways of biomass.
Catalysts 14 00499 g002
Figure 3. The chemical conversion of biomass.
Figure 3. The chemical conversion of biomass.
Catalysts 14 00499 g003
Table 1. The influence of catalysts on the products of biomass conversion.
Table 1. The influence of catalysts on the products of biomass conversion.
CatalystBiomassOperating ConditionsConversion ProductsReference
THRSGFBio-CharBio-OilBio-Gas
K2CO3Empty fruit bunch600 °C10 °C/min200 (N2) mL/min24.3 wt%28.4 wt%46.7 wt%[25]
MgCl2Yellow poplar wood 500 °C _7.7 wt%54.8 wt%37.5 wt%[26]
K3PO4Switchgrass400 °C_1500 (N2) mL/min33.3 wt%42.2 wt%24.5 wt%[27]
MgOPalm empty fruit bunches500 °C25.8 °C/min20–40 (N2) cm3/min25.7 wt%42.3 wt%32.0 wt%[28]
CaOCotton stalk600 °C13 °C/s100 (N2) mL/min24.8 wt%49.7 wt%26.4 wt%[29]
Ca(OH)2Empty fruit bunch600 °C10 °C/min200 (N2) mL/min10.3 wt%42.6 wt%47.1 wt%[25]
ZnOPalm empty fruit bunches500 °C25.8 °C/min20–40 (N2) cm3/min27 wt%44.7 wt%28.2 wt%[28]
TiO2Poplar wood500 °C_120 (N2) mL/min29.7 wt%49.8 wt%20.1 wt%[30]
Activated carbonSwitchgrass400 °C_1500 (N2) mL/min28.7 wt%27.5 wt%43.8 wt%[27]
Bamboo-based bio-charMoso bamboo700 °C_100 (N2) mL/min18.7 wt%17.2 wt%64.1 wt%[31]
KOH-activated pine charPine sawdust450 °C_300 (N2) mL/min28.4 wt%17.0 wt%38.4 wt%[32]
Natural zeolite (clinoptilolite)Switchgrass400 °C_1500 (N2) mL/min30 wt%34.1 wt%35.9 wt%[27]
HZSM-5 (Si/Al = 50)Rape straw550 °C60 °C/min20 (N2) mL/min29.3 wt%27.5 wt%40.7 wt%[33]
Layered zeolites (Si/Al = 35)Microalgae500 °C1000 °C/min_39.2 wt%27.1 wt%34.3 wt%[34]
LaSrNi0.8Fe0.2O3Bamboo sawdust700 °C_300 (N2) mL/min17.5 wt%42.0 wt%40.6 wt%[35]
Al2O3MnCherry pits600 °C10 °C/min100 (N2) mL/min26.2 wt%12.3 wt%48.1 wt%[36]
T = Temperature, HR = Heating Rate, SGF = Sweeping Gas Flow Rate.
Table 2. The influence of catalysts on the preparation of HMF and its derivatives from biomass.
Table 2. The influence of catalysts on the preparation of HMF and its derivatives from biomass.
CatalystReactantConversion ProductsOperating ConditionsYieldReference
TemperatureTime (h)
AlCl3GlucoseHMF220 W microwave radiation0.1674.04%[66]
Reduced-type graphene oxide(rGO)GlucoseHMF200 °C0.57.4%[67]
Ni-rGOGlucoseHMF200 °C0.528.1%[67]
Co-rGOGlucoseHMF200 °C0.517.8%[67]
NiGO-FDGlucoseHMF200 °C195%[67]
UiO-66-SO3H-NH2/PDA@GOGlucoseHMF120 °C255.8%[68]
Υ-Al2O3GlucoseHMF175 °C0.2552%[69]
CrCl3·6H2OGlucoseHMF130 °C0.1671.5%[70]
Co3O4/Co2MnO4HMFFDCA120 °C2498%[71]
Co-SAs/N@CFuroate FDCA260 °C3671.1%[72]
Co-SAs/N@CHMFFDCA85 °C899.5%[72]
Ni(OH)2/SiO2HMFBHMF195 °C598.9%[73]
Ov-WO3@CB[5]HMFBHMF25 °C1>99%[74]
Pd-Ir/CHMFBHMF80 °C2.594.7%[75]
Table 3. The influence of catalysts on the preparation of LA and its derivatives from biomass.
Table 3. The influence of catalysts on the preparation of LA and its derivatives from biomass.
CatalystReactantConversion
Product
Operating
Conditions
YieldReference
TemperatureTime (h)
Zr-HY zeolite catalystsGlucoseLA200 °C33.0%[81]
15-Zr-HYGlucoseLA200 °C35.6%[81]
Carbon foam-supported H5AlW12O40CelluloseLA180 °C460.9%[82]
H3PW12O40CelluloseLA180 °C88.4%[82]
AlCl3CelluloseLA180 °C834.6%[82]
H5AlW12O40CelluloseLA180 °C869.6%[82]
Nb2O5Palm mesocarp fiberLA180 °C-13.2%[83]
H3PW12O40Palm mesocarp fiberLA180 °C-10.5%[83]
HPW-Nb2O5Palm mesocarp fiberLA180 °C-16.4%[83]
Na-BEAGlucoseLA200 °C5100%[84]
CuZ(60%)@HGlucoseLA200 °C545%[85]
Ni1Co1/CLAGVL160 °C382.3%[86]
Ru/ACLAGVL30 °C3.592.5%[87]
Ru@me-HNCLAGVL30 °C3.598.2%[87]
Ru@tk-HNCLAGVL30 °C3.598.1%[87]
SBA-15LAGVL200 °C2<5%[88]
ZrP(15)/SBA-15LAGVL200 °C261.5%[88]
ZrP(15)/SBA-15LAGVL140 °C243.7%[88]
Ov-WO3@CB[5]LAGVL25 °C199%[74]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cai, J.; Wei, L.; Wang, J.; Lin, N.; Li, Y.; Li, F.; Zha, X.; Li, W. Application of Catalysts in the Conversion of Biomass and Its Derivatives. Catalysts 2024, 14, 499. https://doi.org/10.3390/catal14080499

AMA Style

Cai J, Wei L, Wang J, Lin N, Li Y, Li F, Zha X, Li W. Application of Catalysts in the Conversion of Biomass and Its Derivatives. Catalysts. 2024; 14(8):499. https://doi.org/10.3390/catal14080499

Chicago/Turabian Style

Cai, Jixiang, Lianghuan Wei, Jianguo Wang, Ning Lin, Youwen Li, Feixing Li, Xianghao Zha, and Weizun Li. 2024. "Application of Catalysts in the Conversion of Biomass and Its Derivatives" Catalysts 14, no. 8: 499. https://doi.org/10.3390/catal14080499

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

Article Metrics

Back to TopTop