Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Recent advances in biomass valorization through thermochemical processes, bio-oil production and AI strategies: a concise review

Yasmeen Saleh a, Labeeb Ali *ab and Mohammednoor Altarawneh *a
aUnited Arab Emirates University, Department of Chemical and Petroleum Engineering, Sheikh Khalifa bin Zayed Street, Al-Ain 15551, United Arab Emirates. E-mail: labeeb_ali@uaeu.ac.ae
bAbu Dhabi Polytechnic University, Petroleum Engineering Technology Department, Abu Dhabi, 111499, United Arab Emirates. E-mail: mn.altarawneh@uaeu.ac.ae

Received 7th August 2025 , Accepted 21st October 2025

First published on 24th November 2025


Abstract

Biomass valorization through thermochemical pathways is an appropriate option to generate bioenergy, reducing the reliance on petroleum-based fuels and minimizing their adverse environmental impacts. Biomass waste accumulates abundantly worldwide and is typically managed through landfilling and incineration, posing threats to the environment. Alternatively, biomass waste can be converted into valuable products through thermochemical techniques. This review provides an overview on the categories of biomass and the most widely utilized thermochemical pathways for its conversion, including pyrolysis, gasification, liquefaction, and combustion. Additionally, the characteristics and conditions of the processes, configuration of the utilized reactors, the main products, bio-oil catalytic upgrading, advanced upgrading techniques, and commercial-scale thermochemical plants are discussed. Moreover, it summarizes the main findings of the notable studies that investigated the thermal degradation of various biomass types. Furthermore, it highlights the role of artificial intelligence (AI) in forecasting bioenergy production from lignocellulosic biomass. Finally, the main challenges in the thermochemical valorization of biomass, which necessitate further research, are identified.


image file: d5ra05770a-p1.tif

Yasmeen Saleh

Yasmeen Saleh recently received her Master's in Chemical Engineering from the UAE University. Her research focused on thermochemical processes such as pyrolysis to produce renewable fuels and green chemicals, with the aim of advancing sustainable and low-carbon energy solutions.

image file: d5ra05770a-p2.tif

Labeeb Ali

Labeeb Ali holds a PhD in Chemical Engineering from the United Arab Emirates University (UAEU), UAE. He obtained a Master of Science in Chemical Engineering from UAEU. He is currently an Assistant Professor at Abu Dhabi Polytechnic University. With several research and review articles published in high-impact journals, he has made extensive contributions to the field of chemical engineering. His active research areas include thermochemical conversions, biomass valorization, halogenated pollutants, waste management, heterogeneous catalysis, energy conversion, and environmental pollutants, and he is devoted in developing novel materials for energy and environmental-related fields, targeting circularity and sustainability.

image file: d5ra05770a-p3.tif

Mohammednoor Altarawneh

Mohammednoor Altarawneh received his PhD in Computational Physical Chemistry from the University of Newcastle, Australia, in 2008. He is currently a Professor of Chemical Engineering at the United Arab Emirates University. His research focuses on combustion kinetics, heterogeneous catalysis, and formation of halogenated pollutants.


1. Introduction

The tremendous rise in fossil fuel consumption over the past few decades due to the rapid population growth and industrialization has led to the depletion of petroleum-based fuel resources and irrevocable impacts on the environment, particularly the emission of greenhouse gases. As a result, several countries are now actively working to adopt sustainable and renewable energy sources to mitigate the climate change crisis and shift their dependency on conventional fuel resources.1–3 According to the International Energy Agency (IEA), it is expected that the global demand for biofuel will grow by 41 billion liters by 2026.4 Biomass is a term that describes the organic material that originates from plants via the photosynthesis process;2 it is a renewable energy source that has the ability to be converted into biofuel and chemicals through the thermal breakdown of lignocellulosic materials.5 Biomass is a promising alternative to petroleum-based fuels due to its availability and carbon neutrality, spanning diverse resources such as agricultural waste, food processing waste, energy crops, municipal solid waste, and forestry waste.6–8 The aforementioned categories can be transformed into bio-oil, syngas, and biochar through thermal degradation, in addition to valuable products, such as ethanol, furfural, and phenol.9,10 The major thermochemical valorization processes include pyrolysis, gasification, combustion, and liquefaction.11 Combustion is the only process that produces heat, which can be used directly for electricity generation, while the pyrolysis, gasification, and liquefaction techniques demand energy input to yield bioenergy. In addition to the generation of biofuel, the thermal conversion of biomass has the potential to produce valuable chemicals, like plastics, fertilizers, and pharmaceuticals, that have applications in several industries.12–14 The end products of thermochemical conversion depend principally on the treatment process, operating conditions, and biomass feedstock composition.5 Artificial intelligence (AI), which focuses on developing models capable of learning from datasets and making decisions, is widely utilized in various industrial sectors. Biomass thermochemical conversion for bioenergy production is one of the sectors that can employ AI algorithms to enhance its performance and productivity.15

The aim of the present review is to discuss the major biomass categories and the general characteristics of the main thermochemical techniques applied for biomass conversion. Moreover, it provides an overview of the commonly used reactors, the characteristics of the products (bio-oil, pyrolysis gas, and biochar), bio-oil upgrading, and the role of AI in biomass conversion, while also summarizing notable studies on biomass valorization. In addition, the main challenges associated with biomass conversion through thermochemical techniques are discussed.

2. Biomass categories

Biomass is an inclusive term that describes the organic matter produced by the photosynthesis process, existing in the form of plants and microorganisms.1 The solar energy stored in plants is called biomass energy, which can be recovered by thermochemical processes.8 In the photosynthesis process, plants absorb water, CO2, and solar energy from the atmosphere to produce oxygen and monosaccharides such as glucose, which are further converted into secondary products such as polysaccharides, proteins, and lipids. The common categories of biomass are shown in Fig. 1, which include agricultural crop residues, forestry waste, food processing waste, energy crops, municipal waste, and animals waste.6–8 Annually, around 180 billion tons of lignocellulosic biomass are produced from agricultural and forestry residues, while only 8.2 billion tons are reused.16,17 Most of the produced biomass is managed via landfilling or incineration, posing adverse environmental impacts. Instead, lignocellulosic biomass can be valorized into a variety of products such as fuel substitutes, bio-plastics, fertilizers, chemicals, and construction materials.17,18
image file: d5ra05770a-f1.tif
Fig. 1 Major biomass categories and some examples.

2.1. Agricultural waste

Significant quantities of crop residues are produced annually worldwide during harvesting or processing.8 The world's population rapid growth has increased the demands on crop production, further leading to an increase in the generated crops waste.19 These residues comprise straws, stalks, seeds, leaves, husks, etc.,20 which essentially consist of lignocellulosic materials. Annually, about 140 billion tons of agricultural waste are generated but only 40% is reused for fuel production and power generation.17 Rice husk is the most common agricultural waste, which makes up 25% of rice mass.8 About 66% of agricultural waste originates from cereal crop residues. Sugarcane waste, such as stems and leaves, is another major contributor to agricultural waste.18 Accordingly, agricultural residues are the preferred option for bioenergy generation owing to their abundant availability, low CO2 emissions, diverse products, and commercial potential.21,22 These residues have the potential to be recycled into valuable products instead of landfilling or incineration.17

2.2. Food processing waste

The food processing industry produces large amounts of wastes. These processes involve separation of the nutritional portion of raw fruits and vegetables from inedible parts.23 Approximately 38% of food waste is attributed to the food processing industry, where the main contributors to food waste are oil-bearing crops, fruits, and vegetables. As reported by the Food and Agriculture Organization (FAO) of the United Nations, about one third of the fruits and vegetables produced annually is converted into by-products in the food processing industry.24,25 The residues from processing fruits, vegetables, cereals, and oil crops include peels, seeds, shells, stems, and pulp.26

2.3. Energy crops

Energy crops are plants that are grown especially for energy generation, which can be utilized to produce biogas or liquid fuel.27 Oils in the form of fatty acids and lipids are stored in a variety of energy crop seeds, which can be converted to biodiesel and biofuel.5 The crops selected for bioenergy production should have the characteristics such as rapid growth, low cost, ability to adapt to severe soil and weather conditions, capability to offer high yields, low nutrient requirements, and not competing with the food production system.7,28 Examples of energy crops include cardoon, kenaf, sorghum, safflower, poplars, hemp, and rapeseed.7

2.4. Forestry waste

Forestry waste is generated from forests due to logging operations and wood industrial activities such as processing of timber, plywood, and pulpwood. Wood processing activities produce residues comprised of sawdust, barks, branches, split wood, etc..29 In developed countries, wood processing residues are utilized for energy generation at the source of production to avoid wood waste transportation costs. The majority of energy produced from biomass utilizes wood and wood residues.8

2.5. Municipal solid waste

Municipal solid waste (MSW) is the waste disposed daily from households and from commercial activities, which can be organic and inorganic. Organic MSW includes paper, plastic bottles, food residues, textiles, wood, and furfural residues.30,31 Annually, around 1.3 to 1.9 billion tons of solid waste are generated globally, which is expected to reach 3.5 billion tons in 2050 due to the rapid increase in the human population and industrialization.1,32 MSW is typically disposed by landfilling, which could contaminate groundwater, in addition to the requirement of large land area. Nevertheless, the methane produced from landfilling can be used as an energy source.8 Thermal treatment by incineration is another common practice to reduce the solid waste volume and generate heat. However, it poses a threat to the environment due to the emission of pollutants such as nitrogen oxide, sulfur oxide, and fly ash.32

3. Biomass composition

Lignocellulosic biomass is comprised of hemicellulose (15–30%), cellulose (40–60%), and lignin (15–30%), as shown in Fig. 2. The content of these three components varies depending on the type of biomass.3 In addition to the above-mentioned three constituents, biomass contains inorganic ash and minor fractions of extractives such as lipids, sugars, proteins, and starch.33 In plant cells, cellulose forms rigid microfibers, which function as the skeleton of the cells, whereas lignin and hemicellulose pack the inner cell space.3 Cellulose is a linear polysaccharide composed of long chains of glucose units linked by β-1,4-glysodic bonds, which tend to decompose when exposed to high temperatures. The abundance of OH groups in cellulose leads to extensive hydrogen bonding, involving both intra-chain and inter-chain interactions. Intra-chain hydrogen bonds stabilize the glycosidic linkages, while inter-chain hydrogen bonds promote the parallel stacking of cellulose chains.34 Cellulose rapidly degrades during fast pyrolysis as its weak β-1,4-glycosidic bonds cleave under acid or high-temperature conditions, reducing the degree of polymerization and producing furans and levoglucosan.3 Hemicellulose is a heterogeneous polysaccharide consisting of short pyranose and furanose sugar units such as D-xylose, D-mannose, D-glucose, D-galactose, L-arabinose, galacturonic acid, and glucuronic acid. It has a lower degree of polymerization compared to cellulose. Hemicellulose accounts for approximately 15–35% of the dry mass of plants, and its composition varies with plant species.35 The branched structure of hemicellulose accounts for its lower thermal stability compared to cellulose.36 Lignin is a branched polymer primarily composed of three aromatic basic units, i.e., p-coumaryl (H), coniferyl (G), and sinapyl (S) alcohols. These three basic units mainly differ by the number of methoxy groups on their aromatic ring, and their H/G/S proportions largely depend on the biomass species. Lignin linkages are mainly ether bonds (60–70%), followed by carbon–carbon bonds (30–40%), while ester bonds are minimal and mostly found in herbaceous plants.3,37 The complex cross-linked aromatic network of lignin confers higher thermal stability than cellulose and hemicellulose.3,36
image file: d5ra05770a-f2.tif
Fig. 2 Lignocellulosic biomass components.

4. Thermochemical processes

The primary thermochemical processes are pyrolysis, gasification, liquefaction, and combustion,11 where all the aforementioned techniques involve the thermal degradation of the organic components in biomass.1 Thermal conversion treatment can recover the energy stored in biomass into heat in the case of combustion, or into biofuel via pyrolysis, gasification, and liquefaction.37 The choice of thermochemical technique is affected by several factors such as biomass feedstock type, desired end products, process economics, and environmental concerns.1,2 Fig. 3 shows the four major thermochemical processes, their conditions, and target products. In the ternary diagram, as shown in Fig. 4, each corner represents 100% of a specific element (carbon, hydrogen, or oxygen). Char, being richer in carbon, is positioned at the carbon corner. For instance, biomass with a higher hydrogen and carbon content, is located near the carbon-hydrogen side in the diagram. Thermochemical processes decompose biomass components, shifting their positions in the diagram depending on the feedstock composition and the process applied. If biomass undergoes carbonization, its composition shifts toward the carbon corner of the ternary diagram. Conversely, hydrogenation enriches the product in hydrogen, moving it closer to the hydrogen corner.38
image file: d5ra05770a-f3.tif
Fig. 3 Various thermochemical processes and their products.

image file: d5ra05770a-f4.tif
Fig. 4 Ternary diagram of biomass thermochemical conversion, adapted from ref. 248 with permission from Elsevier, Energy Convers. Manag., vol. 304, p. 118222, Copyright 2024.

4.1. Pyrolysis

Pyrolysis is a thermochemical process, which involves the irreversible decomposition of biomass due to the intense heat applied (300–700 °C) in the absence of oxygen to yield non-condensable gases, bio-oil, and char. The obtained products are affected by the process temperature, heating rate, residence time, and biomass nature. There are three main categories of pyrolysis, namely, fast, intermediate, and slow, depending on the residence time and the temperature. Fast pyrolysis has a short residence time and moderate temperature between 450–600 °C. It requires high heat transfer rates, which are attained by grounding the feed biomass. Fast pyrolysis mainly promotes the yield of liquid products (50–70 wt%).5,37,39,40 Alternatively, slow pyrolysis (carbonization) has a longer residence time and lower temperature (400 °C), and it produces approximately equal proportions of gases, bio-oil, and char.37 Pyrolysis occurs in three main stages, as shown in Fig. 5, where the first stage is for moisture removal. Further, the dry biomass undergoes primary pyrolysis, during which bond scission of its lignocellulosic components (hemicellulose, cellulose, and lignin) occurs, producing gases (CO, CO2, H2 and methane), H2O, tar, and primary char. At higher temperatures the primary products undergo further reactions such as dehydration, cracking, polymerization, and gasification, producing secondary products which include gases, bio-oil (phenols, ketones, aldehydes, furans, acids, aromatics, etc.), and secondary char.40–42 Both primary and secondary reactions can occur simultaneously. The produced char catalyzes the cracking of the primary products into lighter gases. Moreover, char gasification leads to the formation of gaseous products, and the polymerization reaction produces secondary char.40,42 The evolved gaseous products yield increases as the temperature increases, which is attributed to the degradation of primary char and tar. In contrast, the char yield decreases as the temperature increases due to the release of more volatiles. The maximum bio-oil yield is attained at approximately 450–500 °C.43–46 The pyrolysis of a diverse range of biomass materials has been reported in the literature, as listed in Table 1.
image file: d5ra05770a-f5.tif
Fig. 5 Different stages of pyrolysis.
Table 1 Studies on the pyrolysis of biomass
Biomass Reactor Catalyst Study purpose Main findings References
Hyacinth biomass (agricultural waste) Solar pyrolysis reactor No catalyst To explore the effect of solar-torrefied biomass samples in improving the solar pyrolysis performance and product quality -Higher torrefaction temperatures increased char (27.4 to 59.4 wt%), while reducing bio-oil (40.2 to 15.3 wt%) and syngas (32.4 to 15.3 wt%) yields 47
-FTIR and GC-MS analyses showed that torrefaction reduces bio-oil acidity, lowering the corrosion potential
Cotton stalk (agricultural waste) Fixed-bed reactor No catalyst To investigate how reaction conditions influence the bio-oil yield from slow pyrolysis of cotton stalk Bio-oil yield peaked at 38.5% at 500 °C with 0.5–1 mm cotton stalk particles, dominated by phenols (>40%). Under the same conditions, high-quality biochar was obtained 48
Orange peel, potato peel, pineapple peel, and corn cob (food processing waste) TGA No catalyst To evaluate the potential of food processing waste-derived bio-oil as a source of biofuel and valuable chemicals Producing transportation fuel fractions requires secondary catalytic treatment to reduce N/O-compounds and enhance pyrolytic oil quality 10
-The produced bio-oil can be utilized as a source of chemicals (e.g. phenols and furans) that have pharmaceutical applications
Sugarcane (agricultural waste) Fixed bed batch reactor No catalyst To investigate the effect of different operating conditions on sugarcane slow pyrolysis to improve the yields of bio-oil, biochar, and evolved gases -The highest bio-oil yield (44.7%) was obtained at 550 °C for a 16 cm height bed at 25 °C min−1 heating rate 49
-The obtained bio-oil requires upgrading to be utilized as a fuel
Oil palm mesocarp fiber and palm frond (agricultural waste) Fixed bed reactor No catalyst To explore how the temperature affects the pyrolysis products formation -The maximum bio-oil yield was obtained at 550 and 600 °C 46
-The obtained bio-oil contained oxygenated and aromatic compounds
Hornbeam shell residues (forest waste) Fixed bed reactor No catalyst Optimizing pyrolysis conditions for enhanced bio-oil and biochar production -The maximum bio-oil yield of 24.67% was reached at 550 °C, and the highest bio-char yield of 40.3% was attained at 400 °C 45
-FTIR analysis revealed the dominance of oxygenated compounds
Bamboo (agricultural waste) TGA and laboratory-scale pyrolysis reactor No catalyst To study how the temperature affects slow pyrolysis behavior and the products yield -The primary product was biochar 50
-The maximum yield of bio-oil was 36.57% obtained at 500 °C
Coconut shell, rice husk, sugarcane bagasse, cotton stalk, wheat straw and palm kernel shell. In addition to cellulose, hemicellulose, and lignin (agricultural waste) Two stages fixed bed reactor 10 wt% nickel-based alumina catalyst Exploring hydrogen and syngas production from different biomass samples utilizing two-stage pyrolysis-catalytic steam reforming -Introducing steam and catalyst into the pyrolysis process enhanced hydrogen generation 51
-The highest hydrogen yield obtained for palm kernels, and the lowest for wheat straw
Miscanthus (agricultural waste) Fluidized bed system for pyrolysis, and fixed bed reactor for bio-oil upgrading Pd/C To perform qualitative and quantitative analyses of the obtained and upgraded bio-oil -The oxygenated and sulfur compounds were reduced in the upgraded bio-oil, and the heating value was increased 52
-The upgraded bio-oil is a suitable substitute for transportation fuel
Wood and wood barks of Ficus religiosa (forest waste) TGA and laboratory-scale fluidized bed reactor No catalyst To investigate the pyrolysis behavior of the feedstock under various operating conditions, in addition to conduct complete product distribution via FTIR-GC/MS analysis -The highest yield for bio-oil of 47.5% obtained at 450 °C 53
-The obtained bio-oil comprised a variety of organics that can be utilized in the chemical industry
Orange pulp waste (food processing waste) TGA No catalyst To provide a detailed analysis for the thermal degradation behavior, kinetics, and evolved gas analysis -Activation energy calculation revealed the complexity of the reactions 54
-The main products are CO, CO2, and H2O, in addition to oxygenated, aliphatic, and aromatic compounds
Rapeseed (energy crop) TGA No catalyst To examine the influence of heating rate on rapeseed slow pyrolysis As the heating rate increased, the maximum DTG curve peak temperature decreased 55


The co-pyrolysis of lignocellulosic biomass and synthetic polymers is an efficient pathway for upgrading bio-oil, owing to the synergistic effect.56 Plastic has a high H/C ratio and low oxygen content, which support the low H/C and high O/C ratios in biomass, respectively. During co-pyrolysis, plastic acts as a hydrogen donor, which improves the quality and the heating value of the obtained bio-oil.57,58 Table 2 lists the studies related to the co-pyrolysis of plastic and biomass.

Table 2 Studies on biomass and plastic co-pyrolysis
Biomass/plastic blend Reactor Purpose of the study Main findings References
Lignin/PP, HDPE and LDPE TGA To explore the potential valorization of these materials via co-pyrolysis Py-GC/MS revealed that lignin-LDPE/HDPE mixtures at 600 °C and a 1[thin space (1/6-em)]:[thin space (1/6-em)]3 ratio achieved the highest compound recovery 59
Almond shell/HDPE TGA To explore the synergistic influence on the pyrolytic oil yield, and to investigate its potential as a bioenergy source Co-pyrolysis reduced oxygenated compounds and enhanced aliphatic formation 60
Pine woodchips/waste-tyre scrap Fixed bed reactor and auger reactor pilot plant To explore the co-pyrolysis of pine woodchips and waste tyre scraps to upgrade the bio-oil quality The acidity, density and oxygen content for the obtained bio-oil decreased 61
Oil palm trunk (OPT)/PP TGA To study the characteristics of the bio-oil obtained from OPT pyrolysis and co-pyrolysis of OPT and PP PP presence improved hydrocarbons production and decreased the contents of phenolics in the co-pyrolysis oil 62
Grape stems and plastic waste (PE, PP, and PS) TGA To explore the thermal behavior, pyrolytic characteristics, and the synergistic effect of the blend co-pyrolysis -At the early stages, softened plastic enveloped the biomass and hindered the release of volatiles 63
-Intermediates generated from biomass breakdown enhanced plastic chain degradation
-Hydrogen radicals released from the plastic enhanced the formation of aromatics
Lignin/plastics (PE, PP, PS, and PC) Semi-batch glass reactor To determine the yields and the composition of the products -The degradation of lignin was reduced, whereas that of plastic enhanced 64
-PC/lignin blend produced the highest phenols
Pine wood sawdust, bamboo, and empty fruit bunch/HDPE and PS TGA To investigate and compare the thermal degradation of different plastic and biomass blends -The degradation of the individual biomass showed one single peak in the DTG curve, while that of the blends showed two peaks 65
Cotton stalk, hazelnut shell, sunflower residue, and arid land plant Euphorbia rigida/PCV and PET TGA Comparison between the thermal degradation behavior of biomass and its blend with plastic -The thermal behavior of biomass and plastic was different owing to their structural differences 66
-Higher activation energies needed for plastic degradation compared to biomass


Several pyrolysis reactors have been utilized for the pyrolysis of biomass such as fluidized bed reactors (FBRs), in which biomass particles are mixed with sand and heated rapidly. The bed is heated by combusting a portion of the produced gas and char. Sand is utilized as an inert bed solid, which provides heat transfer to the biomass feed.67 Fluidized bed reactors are classified into three types, which are circulating fluidized bed (CFBR), bubbling (BFBR), and entrained reactors (EFBR), as shown in Fig. 6, 7, and 9a, respectively.68,69 Fluidized bed reactors are easy to scale up, have a simple design, and can be utilized for fast pyrolysis due to the ease of controlling the residence time by controlling the fluidized gas flow rate. However, FBRs are large and require high construction and operational costs.39,70 CFBRs consist of two interconnected units, one for pyrolysis and one for char combustion, enabling continuous circulation of heated sand. Operating at high gas velocities, they ensure short biomass residence times, effective gas–solid contact, uniform mixing, and improved temperature control. Similarly, BFBRs use an inert gas and sand to fluidize small biomass particles, ensuring high heating rates, uniform temperatures, and efficient continuous operation with high bio-oil yields; however, they require careful control of heat transfer and are sensitive to the biomass type, particle size, and gas velocity. In EFBRs, dried biomass is fed at a controlled rate and carried into the reactor by preheated gas. The particle velocity is determined by the gas flow, and pyrolysis occurs as the biomass passes through the heated reaction zone.68,69 Auger (screw) reactors (Fig. 9b) are tubular, continuous systems where biomass is conveyed by a rotating screw, while heat is supplied through the reactor walls. The screw enhances mixing and contacts with the heated surface, promoting efficient pyrolysis. A key advantage of this design is its compactness and potential portability, enabling on-site biomass conversion, which lowers operational costs by reducing feedstock transportation to biorefineries.39,71 Ablative reactors are another type of pyrolysis reactor in which the biomass particles are subjected to a pressure against a heated plate, causing biomass thermal erosion at the hot contact layer, as depicted in Fig. 8a. This type of reactor is capable of processing the large particle size of biomass, and its efficiency depends on the applied pressure and the plate surface temperature. Ablative pyrolysis is influenced by the heated plate surface area, thus scaling up could be challenging.72 Rotary kiln reactors involve a continuous slow pyrolysis process, in which biomass move by blades installed in the inner rotary tube, as shown in Fig. 8b. The biomass is heated by the flue gas produced from syngas combustion. This technology has high efficiency and stable product formation, and it is the most utilized reactor for municipal solid waste pyrolysis. Rotary kiln pyrolysis technology is applied in China for heating purposes in rural areas.73 Other examples of typical pyrolysis reactors include rotating cone and vertical kiln.39,73


image file: d5ra05770a-f6.tif
Fig. 6 Schematic of a circulating fluidized bed reactor (CFBR).

image file: d5ra05770a-f7.tif
Fig. 7 Schematic of a bubbling fluidized bed reactor (BFBR).

image file: d5ra05770a-f8.tif
Fig. 8 Schematic of (a) ablative pyrolysis reactor and (b) rotary kiln reactor.

image file: d5ra05770a-f9.tif
Fig. 9 Schematic of (a) entrained flow reactor (EFR) and (b) Auger reactor.

Over the past, significant efforts have been devoted to commercializing the pyrolysis process. However, although the reaction mechanisms have been extensively studied, other aspects, such as reactor control systems, still require further research. Biomass pyrolysis faces challenges in sustaining continuous operation and producing high-quality bio-oil. Although catalysts and hydrotreatment have been used to improve the quality, rapid catalyst deactivation and high costs have hindered its commercialization. In this case, several companies are pursuing commercialization, and a number of moderate-scale demonstration and commercial plants has already been established.74,75 Table 3 lists some of the commercial pyrolysis plants.

Table 3 Commercial biomass pyrolysis units located worldwide76,77.a
Country Reactor Biomass Capacity Status
a ML: million liters.
Brazil CFBR Forest residues 83 ML year−1 Under development (in 2023)
Canada CFBR Wood residues 38 ML/year Active since 2018
Canada CFBR Mill and forest residues 11 ML year−1 Active since 2006
Canada Auger reactor Wood residues 6 ML year−1 Under development (in 2023)
Finland Rotating cone Sawdust and wood residue 20 ML year−1 Active since 2020
France Cellulosic biomass 0.8 ML year−1 Under development (in 2023)
The Netherlands Rotating cone Wood biomass 20 ML year−1 Active since 2015
Sweden Rotating cone Sawdust 21 ML year−1 Active since 2021


4.2. Gasification

The gasification process involves converting solid and liquid biomass into gaseous products, usually syngas, along with unconverted char and ash, by applying high temperatures.8,78,79 This is a crucial process for green hydrogen production, which can be applied in fuel cell vehicles, power generation, and industrial processes, helping to reduce emissions.80 The gasification process occurs in the presence of gasifying agents, which can be air, oxygen, steam, and their combination, where each agent yields different compositions and heating values for the obtained syngas. Gasifying agents react with carbon and heavy hydrocarbon, producing lower molecular weight gases.81 Heat is provided either through partial combustion of the products (auto-thermal) or from an external source (allo-thermal). As depicted in Fig. 10, the gasifying process involves four main reactions, including drying (below 150 °C), pyrolysis (150–700 °C), oxidation (700–1500 °C), and reduction (800–1100 °C).82 In the pyrolysis reaction, as discussed earlier, biomass decomposes, releasing gaseous products, tar, and char. The pyrolysis products undergo partial oxidation, producing CO, CO2, and water, while also releasing the thermal energy needed for other endothermic reactions (drying and pyrolysis). Reduction reactions incorporate all the pyrolysis products to yield the final syngas.79,83 The properties of the products depend on the biomass feed type, gasifying agent, catalyst, and temperature.83 Gasifier reactors are classified according to the gasifying agents, heat source, and gasifier pressure and design, as shown in Fig. 11. Among them, fixed bed gasifiers are the simplest in design, which consist of a bed of solid fuel particles in which the gasifying agent moves through. They are classified into updraft and downdraft reactors. In updraft gasifiers, the gasifying agent and syngas move up (Fig. 12a), whereas in downdraft gasifiers, the syngas moves down (Fig. 12b). Fixed bed gasifiers can handle biomass with high humidity and different sizes and have a high solid residence time, high thermal efficiency, and high carbon conversion. However, fixed bed gasifiers require the treated feed to be homogenous and have the same characteristics.79,84,85 Entrained flow gasifiers are another type of reactor, in which the gasifying agent and biomass are fed in co-current, as shown in Fig. 12c. This type of gasifier operates at high temperature (1200–1500 °C) and requires fine feedstock particles for more efficient heat transfer.79,81,86 Usually, torrefaction pre-treatment of the feedstock is required to reduce the moisture content and minimize the bulk density.87,88 Torrefaction is a thermochemical pretreatment used to upgrade raw solid biomass for gasification. Torrefaction involves three main reactions including depolymerization, devolatilization, and decomposition.89 This process, which improves fuel quality by removing moisture and breaking down rigid fibrous structures, depends on the reaction temperature and type of reaction medium. It is classified into dry (200–300 °C) and wet (180–260 °C). Wet torrefaction is suitable for wet biomass, while dry torrefaction is applied to solid biomass under either oxidative or non-oxidative conditions. Dry torrefaction is commonly used to improve biomass by increasing its energy density and reducing its moisture, oxygen, and volatile content.90,91 In many cases, torrefaction is combined with pelletization to increase the bulk density of the material.79 Abdulyekeen et al.92 evaluated torrefied wood and garden waste in a helical screw fluidized bed reactor at 250–300 °C by analyzing their grindability, thermal, and structural properties. The results showed that temperature had the greatest impact, with higher temperatures reducing the volatile matter, oxygen, and hydrogen, while increasing the fixed carbon, calorific value, and carbon content. Another recent study by Silveira et al.93 explored torrefaction (225–275 °C, 60 min) to increase the energy density of a six-species waste wood blend. TGA was used to examine the combustion behavior, and the emissions were numerically assessed. Torrefaction reduced the H/C ratio from 1.87 to 1.05 and the O/C ratio from 0.70 to 0.47, improving the thermal stability and altering the ignition dynamics, thereby enhancing the overall efficiency of the biofuel. Table 4 lists the studies pertaining to the gasification of various biomass materials, and Table 5 presents some of the industrial applications of biomass gasification.
image file: d5ra05770a-f10.tif
Fig. 10 Main reactions in the gasification of biomass.

image file: d5ra05770a-f11.tif
Fig. 11 Classification of gasifier reactors.

image file: d5ra05770a-f12.tif
Fig. 12 Schematic of (a) updraft gasifier, (b) downdraft gasifier, and (c) entrained flow gasifier.
Table 4 Studies related to biomass gasification
Biomass Gasifier type Study purpose Main findings References
Oil palm wastes (oil palm fronds, empty fruit bunches, oil palm trunks, and palm kernel shells) (agricultural waste) Aspen plus simulation To explore hydrogen production from four oil palm wastes via biomass gasification, in addition to techno-economic, and environmental analysis -The highest hydrogen yield was obtained from oil palm trunks 80
-Hydrogen production from oil palm waste gasification has a longer payback period but lower environmental impact than conventional steam methane reforming
Food waste Batch reactor To explore the potential of supercritical water gasification technology for energy generation and the feasibility of hydrogen production in Bangladesh Electricity generation from food waste in Bangladesh could increase from roughly 489 MW in 2023 to about 2042 MW by 2042 94
Coconut waste (food processing waste) Gasification reactor To explore how NaCl concentration, reaction time, and temperature affect hydrogen production from coconut oil meal gasification Higher gas yields and H2 selectivity were achieved at 550 °C, 75 min reaction time, and with 10% NaCl addition 95
Oak and corn stalk (agricultural waste) Fixed-bed reactor To analyze the syngas, tar, and biochar generated from oak and corn stalks under varying temperatures using electric gasification and microwave gasification -Microwave gasification produced 3–7% more syngas than electric gasification 96
-Microwave gasification yielded 3–6% more biochar than electric gasification, while electric gasification generated more tar in the 600–800 °C range
Wood Solar furnace To explore the potential of syngas generation from wood utilizing prototype solar reactor The prototype solar reactor was reliable for the gasification of both biomass and waste 97
Rice husk pellet (agricultural waste) Bubbling fluidized bed To explore the advantages of combining torrefaction and gasification processes Torrefaction process contributed in increasing the heating value of rice husk pellet. Conversely, the electric energy demand for the process reduced the process efficiency 98
Pine sawdust (forest waste) Fluidized bed gasifier To explore the feasibility of syngas generation from catalytic gasification of biomass High H2/CO syngas was obtained which can be utilized in methanol production 99
Cellulose as a model compound Fluidized bed gasifier To investigate cellulose gasification utilizing Rh/CeO2/SiO2 catalyst in a fluidized-bed reactor at low temperature Utilizing Rh/CeO2/SiO2 as a catalyst introduced a novel approach for hydrogen and syngas generation at low temperature 100
Municipal sludge Downdraft fixed bed gasifier To explore the municipal sludge gasification based on the sludge characteristics, and to study the produced solid residues properties -The high volatile organics in the sludge indicated that the sludge can be valorized via gasification 101
-Ash content was higher than the common biomass
-The organic functional groups determined by the FTIR analysis in the char were quite similar with those in sludge
Manure (animal waste), sewage sludge and date pits (food processing waste) Fixed bed gasifier To study the effect of pressure, temperature, and ratio of oxygen on the composition and the heating value for the syngas -The properties of the obtained syngas can be controlled by optimizing the biomass feedstock mixing 102
-Changing the operating conditions affected the syngas composition and LHV
Pine sawdust agricultural waste Pyrolysis-gasification coupled reactor To investigate the production of hydrogen-rich syngas utilizing a pyrolysis-gasification coupled reactor -CO2 was the highest in the pyrolysis part, whereas H2 was the highest in gasification part 103
-The main carbon conversion occurred in the pyrolysis part
-The coupled reactor produced less tar
Wood, paper, municipal solid waste, and sawdust (forest waste and MSW) Downdraft gasifier Using different gasification agents (air, oxygen, oxygen-enriched air, and steam) and comparing the obtained products -The highest hydrogen production was associated with steam. Conversely, the lowest was associated with air 104
-Increasing the gasification temperature decreased hydrogen production and reduced the syngas calorific value for all the four agents
-Increasing the biomass feeding rate adversely affected the hydrogen production and energy and exergy efficiencies
Pine sawdust and HDPE blend. Agricultural waste and plastic Fixed bed reactor To explore the efficiency of Ni-Fe@nanofibers/porous carbon catalyst for syngas generation through pyrolysis-gasification process for the blend The optimum biomass-plastic ratio was 1/2, and the best syngas quality was obtained at a gasification temperature of 700 °C 105


Table 5 Industrial application of biomass gasification78,106
Country Reactor type Biomass type Product
Austria Wood chips 2 MW electrical power and 4 MW heat
Denmark Circulating fluidized bed (CFB) 6 MW electrical power and 12 MW heat
Finland Entrained flow gasifier (EF) Biomass and sludges, RDF 10–200 MW heat
Canada (Toronto MSW plant) Municipal solid waste Electricity
Netherlands (Americentrale fuel gas plant) Demolition wood Fuel gas
Germany (Fendoteoc gasification plant) Municipal solid waste Electricity


4.3. Liquefaction

Liquefaction is another thermochemical process that aims to produce bio-oil, in addition to fractions of gaseous product and solid residues by applying a relatively low temperature (250–400 °C) and high pressure (5–20 MPa).107,108 The major difference between liquefaction and other thermochemical techniques (pyrolysis and gasification) is the utilization of solvents such as methanol, ethanol, and water as the reaction medium during the liquefaction process.108 Organic solvents act as hydrogen donors, which improve the bio-oil yield with a low oxygen content. Hydrothermal liquefaction (HTL) takes place under hot pressurized water conditions. The HTL process is comprised of three main steps; firstly, depolymerization of the macromolecules into shorter chains. Subsequently, dehydration and decarboxylation reactions convert the oxygen into water and CO2. The last step involves repolymerization of the small monomers to form coke, which occurs in the case of hydrogen deficiency.5,109,110 The formed aqueous phase can be recycled to the liquefaction reactor, which will reduce the water demand, as shown in Fig. 13, whereas bio-crude oil will require upgrading for further uses.5,109 The high solvent cost and the environmental concerns related to solvent recycling are the main disadvantages of utilizing organic solvents.108,111 The liquefaction process can handle wet feedstock, which eliminates the requirement of the feed drying process in the case of using water as the solvent.108,112 The liquefaction process is affected by chemical factors including biomass type and size, catalyst, solvent, and physical factors such as pressure, temperature, heating rate, and residence time.107 Homogenous catalysts (e.g. organic and inorganic acids) are typically used to reduce char formation.107 Moreira-Mendoza et al.113 studied HTL of agro-waste, microalgae, and defatted seeds, finding that lipid-rich feedstocks gave higher biocrude yields (∼65%) and energy content, while lignocellulosic biomass produced only 10–20%. Using Ni-Pt/Al2O3 or Pd-C/Al2O3 with glycerol enhanced deoxygenation and improved the biocrude quality. Another recent study by Awadakkam et al.114 investigated HTL of Canadian spruce and poplar, identifying them as promising feedstocks. Using an ethanol–water co-solvent with K2CO3 improved the biocrude yield, reduced the oxygen content, and increased the energy density. The co-solvent also acted as an in situ hydrogen donor, minimizing the external hydrogen demand and enhancing the process efficiency. A similar study by Ocampo et al.115 investigated HTL of microalgal biomass using a water-acetone mixture, achieving ∼60 wt% biocrude yield at 5 wt% acetone. This process is economically favorable but produces biocrude with high nitrogen (23.68%) and oxygen (10.68%) contents, requiring further refining.
image file: d5ra05770a-f13.tif
Fig. 13 Typical biomass hydrothermal liquefaction process.

Most HTL research has been carried out in batch reactors, which are simple, robust, and capable of operating under a range of reaction conditions. However, although they are versatile for experimentation, batch reactors are unsuitable for industrial-scale biofuel production. Batch HTL operates intermittently, limiting the efficiency and scalability, whereas continuous HTL offers a more efficient and controllable pathway for bio-oil production.116 A continuous stirred tank reactor (CSTR) uses an impeller and electric heater to ensure thorough mixing and maintain the desired reaction temperature, respectively, enhancing the heat transfer compared to a plug flow reactor (PFR). Although a CSTR alone provides a lower biomass to bio-oil conversion due to its shorter residence times, combining it in series with a PFR extends the residence time, increases the bio-oil yield, and minimizes plugging issues with lignocellulosic feedstocks.116–118 This CSTR-PFR configuration, as depicted in Fig. 14, demonstrates strong potential for HTL processes.117 Table 6 lists the studies related to biomass liquefaction.


image file: d5ra05770a-f14.tif
Fig. 14 CSTR-PFR configuration in HTL, adapted from ref. 116 with permission from Elsevier, Energy Convers. Manag., vol. 302, p. 118093, Copyright 2024.
Table 6 Studies related to biomass liquefaction
Biomass Solvent Study purpose Main findings References
Softwood trees residues (forest waste) Water To investigate a novel method for biofuel production by combining HTL with pipeline hydro-transport of biomass -The biocrude has a low oxygen and water content and high heating value, but high acidity and carbon residue indicate the need for catalytic upgrading 119
-Integrating pipeline hydro-transport with HTL is technically feasible for biofuel production, supporting its future commercial implementation
Oil processing residues, agricultural residues, forest residues, and food waste Water To examine biocrude production from different feedstocks via HTL, highlighting how biomass composition influences the HTL process -Mustard meal produced the highest biocrude yield, followed by canola meal, under the optimized conditions of 320 °C, 30 min residence time, and a 5[thin space (1/6-em)]:[thin space (1/6-em)]1 water-to-biomass ratio (wt/wt) 120
-The optimum 1[thin space (1/6-em)]:[thin space (1/6-em)]3 canola to mustard meal ratio yielded 41.9 wt% biocrude with a low oxygen content
Wood bark (forest waste) Water To compare the liquid composition produced via slow pyrolysis and HTL processes for different wood types -HTL liquids contained mainly water, acetic acid and catechol 121
-Slow pyrolysis liquids composed of a variety of phenolic compounds and furfurals
-Similar concentrations of aromatics and phenols were found for HTL and slow pyrolysis
Bamboo (agricultural waste) Phenol, ethylene glycol (EG), and ethylene carbonate (EC) To explore the effect of different organic solvents, liquid ratio, and reaction time on liquefaction yield and products -Phenol is the best solvent for bamboo liquefaction 122
-Similar low boiling point products were identified by GC-MS, irrespective of the solvent type
Green waste mixture of straw, wood, and grass (agricultural waste) Tetralin To investigate the products of green waste liquefaction using Raney Ni catalyst and tetralin as the solvent -The catalyst enhanced methane formation 111
-The catalyst lowered the oxygen content and increased the yield of n-hexane
Pinewood (forest waste) Water, ethanol and acetone To investigate the effect of different solvents on the biomass liquefaction process -Acetone as the solvent resulted in the highest conversion rate 123
-The product distribution was strongly affected by the solvent type
Silver birch (forest waste) Water To study the liquefaction of silver birch wood in liquid water using sodium carbonate as the catalyst and to analyze the products via FTIR and GC-MS -53.3% heavy oil yield was obtained at a reaction temperature of 380 °C 124
-FTIR and GC/MS analysis revealed the formation of aldehyde, ketone, hydroxybenzene and ester
Onopordum heteracanthum stalks (agricultural waste) Methanol, ethanol and acetone To explore the liquefaction of Onopordum heteracanthum stalks with catalysts (KOH and ZnCl2) and without catalyst at three temperatures and to explore the potential of bio-oil production -In non-catalytic liquefaction, the best operating conditions were with acetone at 563 K 125
-ZnCl2 catalyst was found to be more effective than KOH
Cunninghamia lanceolata (forest waste) Water To study the direct liquefaction of Cunninghamia lanceolata The highest heavy oil yield occurred at 280–360 °C 126
Pinus banksiana (forest waste) Ethanol To investigate the liquefaction of Jack pine wood powder in H2 environment using sub/super-critical solution of ethanol and iron-based catalysts (FeS or FeSO4) -Increasing the residence time, H2 pressure, and reaction temperature enhanced the oil yield 127
-(FeS and FeSO4) catalysts improved the oil production
-GC-MS revealed that phenols were the dominant in the bio-oil


The demonstration and commercial plants for HTL are listed in Table 7.76,77

Table 7 Biomass HTL plants76,77 a
Country Plant scale Biomass Capacity Status
a ML: million liters.
Canada Commercial Forestry waste 8 ML per year (bio-oil) Under development (in 2023)
Australia Demonstration Post-consumer waste 1.6 ML per year (bio-oil) Active since 2012
Canada Demonstration Sewage sludge 2 dry tonnes per day Under development (in 2023)
Denmark Demonstration Wastewater sludge 5 ML per year (bio-oil) Under development (in 2023)
India Demonstration Food waste, sludge, and algae 80 L per day (bio-oil) Active since 2016
Norway Demonstration Forest waste 1.5 ML per year (bio-oil) Active since 2021
Turkey Demonstration Different wastes 8.7 ML per year (bio-oil) Active since 2011


4.4. Combustion

The combustion process is comprised of three primary stages, i.e., drying, evolution of volatiles, and char oxidation. The combustion of the released volatiles is the primary source of the generated heat, which is utilized to produce steam for electricity generation. Char is further oxidized by O2, CO2 and H2O and only ash remains.37,128 Biomass combustion is the most common thermochemical process due to is low cost and processing speed. However, the complexity of biomass and the differences in its chemical structure pose challenges in burner design and control of the operating conditions.129 Moreover, the high content of moisture and volatiles in biomass in comparison to coal complicates the combustion performance.130 Table 8 summarizes the studies on biomass combustion and co-combustion with coal.
Table 8 Studies related to biomass combustion and co-combustion
Biomass type Burner/rector type Study purpose Main findings References
Rice straw and wheat straw Fixed bed combustor To study the combustion behavior of rice straw and wheat straw biomass under different operating conditions It was found that a high air temperature reduced the time required for the volatiles and char to burn 131
Olive residue, pine sawdust and sugarcane bagasse Drop tube furnace To study biomass combustion under both oxy-fuel and air conditions and to compare with coal combustion characteristics -Biomass flame was less sooty than that of coal 132
-Changing from air to oxy-fuel hindered the combustion intensity
Spruce bark, short rotation coppice (SRC) poplar, Danish straw, and torrefied wood Single particle reactor (SPR) To investigate the NO release during char combustion NO reduced inside the char particles got released as the conversion progressed 133
Lignite, vine pruning, olive tree pruning, corn stalk, and poultry litters TGA To investigate the potential of agricultural and animal residues as fuels for co-combustion with lignite -Torrefaction process enhanced the char reactivity 134
-Mixing torrefied biomass with lignite decreased the ignition temperature compared to individual lignite
Corncob, hardwood, and bituminous coal TGA To study the effect of heating rates, mixing ratios, and biomass type on combustion characteristics -Biomass showed better combustion behavior than coal 135
-Increasing the heating rates improved combustion performance
Pine, eucalyptus, and willow Natural gas burner To find correlations between particle mass and the combustion behavior Derivation of empirical formulas for the duration of each combustion stage 136
Palm kernel shell Plug flow reactor To investigate biomass combustion properties at elevated temperature and high heating rate that reflect the real furnace environment -Char combustion required a higher residence time in comparison to drying and devolatilization stages 130
-Particle size showed the major effect on char oxidation, while it was less impactful at drying stage
Beech wood Entrained-flow reactor To formulate a model representing the combustion of woody biomass, and to assess its accuracy The developed model can predict various parameters during biomass combustion 137
Naomaohu coal and pine sawdust TGA To investigate the mass decay behavior and the volatile products of biomass and coal and their blend combustion under inert and air conditions -Blending pine dust with coal decreased the ignition time and NO emissions 138
-Oxygen level in pine dust contributed to C[double bond, length as m-dash]O functional group formation
-The co-combustion exhibited a synergistic behavior under air atmosphere
Peanut shells and coal gangue TGA -To predict the co-combustion characteristics of coal gangue and peanut shell using ANN model and compare it with experimental results -The activation energies of coal and peanut shell co-combustion were less than those for individual components 139
-To investigate the thermodynamic, kinetic properties, and evolved gases via FTIR -Combustion products identified by FTIR were H2O, CO2, CO, CH4, C[double bond, length as m-dash]O, phenols, C–O and NH3
-ANN20 was the most accurate model to predict CG and PS co-combustion


In the past, biomass combustion was not considered an energy source for power plants owing to its low heating value compared to fossil fuels and its inconsistent composition.140,141 However, fossil fuel depletion and escalating costs, in addition to the emerging environmental concerns have led to an interest in utilizing biomass for energy generation via co-combustion with coal.141 Many existing large-scale power plants utilize biomass as a feedstock, as enlisted in Table 9. Fluidized bed combustors (FBC) are optimal technology for large-scale power generation plants, given that they have the ability to handle different feed sizes, moisture contents, shapes, feedstock pulverization is not required, also it have a uniform temperature distribution and high heat transfer coefficient between the bed and heat exchanger.141–143 However, there are some disadvantages such as the requirement of an efficient solid–gas separation system, boiler erosion, and agglomeration of the bed for some types of herbaceous biomass.142 A typical circulating fluidized bed combustion system (CFBC) is shown in Fig. 15.142,144

Table 9 Large-scale power plants utilizing biomass141,145,146
Power plant location Biomass Combustion system Capacity (MW)
Power plant at Port Talbot in Wales/UK Wood CFB 350
Alholmens Kraft power plant/Finland Pulp, paper and timber residues CFB 265
Polaniec biomass power plant/Poland Tree-farming and agricultural by-products CFB 205
Power plant in Florida/USA Sugar cane fibre (bagasse) and recycled urban wood 140



image file: d5ra05770a-f15.tif
Fig. 15 Schematic of a typical CFBC system.

5. Comparison between thermochemical processes

All the previously discussed thermochemical conversion technologies have inherent limitations. They often produce lower-quality biofuels, involve high production and energy costs, and require extensive gas purification.147 Accordingly, understanding the advantages and limitations of each technology is essential for evaluating their performance, commercialization potential, and techno-economic feasibility.5 Unlike combustion, which mainly produces heat and electricity, pyrolysis offers a flexible route for liquid bio-oils and chemicals, where handling diverse feedstocks and moderate scales makes pyrolysis especially suitable for rural applications, providing opportunities for local biofuel generation and flexible energy solutions. However, it faces challenges such as poor bio-oil quality, limited catalyst life, scale-up issues, and the need for extensive upgrading.148 Conversely, gasification produces syngas that can be directly used as fuel or converted into valuable products, making it a versatile thermochemical conversion method. It also offers opportunities for large-scale operations and significant growth potential. However, gasification faces challenges, including high operating and maintenance costs, difficulties in recycling homogeneous and carbon-based catalysts, and high overall costs. HTL is a high energy efficiency method for converting wet biomass into bio-oil, making it particularly well suited for biofuel production. It eliminates the need for drying, and operates under relatively mild conditions, producing bio-oil with lower oxygen and moisture contents than pyrolysis oil. However, HTL faces challenges such as potential reactor corrosion, salt deposition, difficulties in recycling homogeneous or carbon-based catalysts, and variability in bio-oil properties depending on the feedstock and production conditions.5 Table 10 summarizes the major advantages and disadvantages of thermochemical biomass conversion methods.5,149
Table 10 Advantages and disadvantages of thermochemical processes
Thermochemical method Advantages Disadvantages
Pyrolysis -Flexible and scalable process suitable for diverse biomass feedstocks -High costs associated with pre-drying biomass with high moisture content
-Operates at low pressure with minimal feedstock pre-treatment -Char formation can cause abrasion and erosion of equipment
-Produces multiple valuable products: biochar, bio-oil, syngas, and hydrogen -Syngas often requires extensive cleaning before use
  -Significant capital investment is needed
  -Catalyst deactivation due to char deposition
  -Feasibility is established only in large-scale plants
Gasification -Achieves maximum biomass conversion -Suitable for biomass with low moisture content; high-moisture feedstocks require pre-drying
-Produces syngas and biochar -Technology is complex with high capital and operating costs
-Capable of high hydrogen yields -Syngas requires extensive cleaning to remove tar, char, and ash
-High efficiency even at small scales -Feedstocks like manure can release H2S, necessitating additional gas cleaning
-Flexible and scalable in capacity  
-Can handle diverse biomass feedstocks efficiently to produce fuel gas  
Hydrothermal liquefaction -Suitable for high moisture content biomass, thus drying is not required -A high-pressure pump is needed to handle highly concentrated biomass feedstock and prevent reactor clogging
-The product contains less oxygen and moisture than pyrolysis bio-oil -Poor design may cause pipeline or reactor failure under high temperature and pressure
  -Corrosion and salt deposition possibility
Combustion -Technology is simple and widely available -Suitable for low-moisture biomass; high-moisture feedstocks require costly pre-drying, reducing efficiency
-Low operating costs -Generates by-products (soot, ash, NOx, CO, CO2) with potential environmental impacts
-Minimal biomass pre-treatment required -Flue gas cleaning is necessary


6. Artificial intelligence (AI) in biomass conversion

Artificial intelligence (AI) is a sophisticated approach applied widely in various disciplines. It has the advantages of avoiding the lengthy experimentation, rigorous calculations, and high cost. The basic concept of AI is to analyze data, pattern derivation, and accordingly generate predictions.150 AI models for biomass thermochemical processes are important tools for research and industry, as they can establish a connection among biomass composition, operating conditions, and biomass utilization.151 AI algorithms depend on large datasets for model development, which achieves accurate results and ensures the formation of products with the desired properties by optimizing the operating conditions.152 AI models have been effectively utilized to predict the outputs of biomass thermochemical conversion, as listed in Table 12.

The application of AI in the biomass feedstock stage is valuable for predicting the yields from climate and cultivation variables, detecting anomalies, and integrating with control systems to improve the quality and efficiency. In the bioenergy production stage, processes such as pyrolysis and hydrothermal liquefaction are energy intensive, but their dependence on controllable reaction parameters makes them suitable for AI-based optimization to enhance the bioenergy yield and overall process efficiency.153 Schmidt et al.154 applied Bayesian optimization to tune the hyperparameters of various AI models for predicting biomass gasification gases. The results showed that XGBoost delivered the highest accuracy and robustness. In the biofuel application stage, AI research has focused on linking biofuel properties such as cetane number, density, viscosity, and calorific value to engine performance, aiming to enhance the combustion efficiency and lower emissions.153 For instance, Yatim et al.155 proved that the ANN model is suitable for predicting the biodiesel density, offering guidance for engine developers on operating conditions and biodiesel selection. In addition to the contribution of AI to improving the biomass yields, it also reduces waste by predicting by-product volumes and compositions, enabling their reuse as feedstock in other processes. This supports circular economy practices by converting waste into valuable resources.156 Moreover, AI-driven demand forecasting is transforming green energy management in manufacturing by accurately predicting energy needs and aligning them with production schedules. This approach minimizes waste, reduces emissions, and strengthens the sustainability.157

The most frequently used machine learning (ML) models for thermochemical biomass conversion include artificial neural networks (ANNs), support vector machines (SVMs), decision trees (DTs), and random forests (RFs). ANNs have been the most prominent method for pyrolysis process modeling due to their ability to learn highly nonlinear relationships between inputs and outputs, making them well suited for capturing the complexity of the pyrolysis process. Likewise, in gasification, ANNs have also been the most widely used ML method.158,159 SVMs can handle classification and regression with low generalization error and minimal computing cost, using kernel functions for complex, nonlinear relationships. DTs recursively split data into a tree structure of root, decision, and leaf nodes, offering easy interpretation and handling of missing data but are prone to overfitting and poor generalization. RFs are ensembles of decision trees that improve accuracy and reduce overfitting by training multiple trees on bootstrapped datasets and random feature subsets, though they can be slow for real-time predictions with large numbers of trees. Other ML methods, such as extreme gradient boosting (XGBoost) regression and adaptive network-based fuzzy inference system (ANFIS), are also employed in biomass conversion.159,160 Ensemble methods such as gradient boosting regression (GBR) improve the prediction reliability by combining multiple models, each capturing different aspects of the data. Madadi et al.161 found that GBR was the most effective in optimizing the biphasic pretreatment conditions for lignocellulosic biomass fractionation. Another recent study by Qiao et al.162 showed that CatBoost Regressor outperformed the other seven evaluated models in optimizing the conversion of cellulose into 5-hydroxymethylfurfural. Table 11 lists the advantages and limitations of some ML models.

Table 11 Advantages and limitations of ML models used in thermochemical biomass conversion158,159
Model Advantages Limitations
ANN Ability to learn highly nonlinear relationships between inputs and outputs -Low precision, preprocessing needs, and overfitting risk
-Incapable of explaining their reasoning or providing a basis for their decisions
SVM Resist overfitting, have low generalization error, and perform efficiently on small datasets Highly sensitive to kernel choice and parameter tuning
DT Easy to interpret and handle missing or duplicate data Prone to overfitting and poor generalization
RF Handle large datasets and reduces overfitting Too many trees slow real-time predictions despite fast training
XGBoost -Precise loss calculation using first and second order derivatives -Can overfit small datasets if not properly tuned
-Reduce overfitting through regularization -Computationally expensive, particularly on large datasets or with complex models


Table 12 Studies on AI models applied in biomass thermochemical conversion
Process Biomass Input Output Algorithm References
Co-pyrolysis 52 distinct lignocellulosic biomasses and 55 plastic types Process and feedstock characteristics Biochar and liquid yields CatB and XGB 163
Microwave-assisted co-pyrolysis A variety of biomass feedstocks such as lignin, rice straw, Chlorella vulgaris, and plastic feedstock (LDPE, PS, and PVC) Proximate and ultimate analyses of biomass and plastic, heating rate, pyrolysis temperature, proportion of biomass and plastic in the mixture Biochar, bio-oil, and biogas yields XGBoost 164
Pyrolysis 15 types of lignocellulosic biomass, 12 types of plants, and 6 types of algae -Ultimate and proximate analyses Bio-oil yield -Random Forest 165
-Heating rate, particle size, highest treatment temperature, and N2 flow rate -Decision tree
  -Support vector machine
  -Multi linear regression
Pyrolysis Herbaceous plants, lignocellulosic biomass, sewage sludge, and animal waste -Ultimate, proximate, and structural analyses of biomass Biochar yield ANN and metaheuristic algorithms 166
-Pyrolysis conditions
Torrefaction Agricultural residues, forestry residences, and energy crops -Ultimate and proximate analyses The yield of solid products Gradient tree boosting was the most accurate 167
-Torrefaction conditions (temperature, residence time, (CO2, CO, N2 fractions in the reacting gas)
Hydrothermal liquefaction Pine The carbon contents, reaction temperature, and biomass concentration Biocrudes heating values Support vector machine (SVM) 168
Acid-catalyzed liquefaction Lignocellulosic biomass Biomass composition, particle size, solvents, catalyst, and liquefaction conditions Bio-polyols yield and heating value GBR 169
Gasification 20 different biomass samples (agricultural and herbaceous waste) Temperature, steam and fuel flow rates, and the content of H, O and C Syngas exergy value ANN and Aspen Plus® model 170
Gasification Sugarcane bagasse Biomass characteristics and gasification conditions Syngas yield XGBoost 171
Gasification Lignocellulosic biomass Laboratory and pilot plant experimental data Gasification products XGBoost 154


7. Bio-oil, biochar and pyrolysis gas characteristics

Bio-oil is a mixture of various molecular weight components produced from the thermal degradation of lignocellulosic materials. It features a dark-brown colored liquid and viscous texture consisting mainly of carbon, oxygen, and hydrogen. The major categories of bio-oil compounds encompass phenols, aromatics, aldehydes, ketones, esters, carboxylic acids, alkanes, and nitrogen-bearing compounds.172 Typically, bio-oil consists of water (20–30%), organics (20–30%), water-soluble pyrolytic lignin (28–36%), and water-insoluble pyrolytic lignin (15–23%).173–175 In addition to its potential as a biofuel source, bio-oil can be utilized as a feedstock for manufacturing plastics, fertilizers, pesticides, pharmaceuticals, carbonaceous materials, etc..12–14 A recent study by Ucella-Filho et al.176 investigated the antiviral potential of Citrus sinensis bio-oil produced at different heating rates. One sample exhibited anti-SARS-CoV-2 activity, and the findings offer valuable guidance for future research on developing bioactive products from bio-oil. Another study by Abd-Elnabi et al.177 reported that olive cake bio-oil contains bioactive compounds with potential as biopesticides. An insect bioassay against aphids, conducted using the slide dip method, demonstrated significant mortality in these sucking pests. Bio-oil has the advantages of carbon neutrality and lower SOx emissions. However, the direct application of bio-oil as a fuel substitute is hampered due to its high viscosity, high moisture content, low calorific value, elevated oxygen content (10–40%), and acidity (corrosiveness). Biochar is the solid material that remains after the thermochemical degradation of biomass in the absence or lack of oxygen (pyrolysis), which basically consists of carbon, in addition to hydrogen, oxygen, and ash. The composition of biochar varies depending on the biomass feedstock, pyrolysis temperature, and heating rate. Lignin-rich biomass yields a high fraction of biochar.178,179 As the pyrolysis temperature exceeds 400 °C, the surface area of the biochar begins to increase up about 900 °C, which is attributed to the release of volatiles inducing pores formation.180,181 Biochar has several applications, for instance, heat generation, soil remediation, adsorbent for air and water pollutants owing to its porous structure, flue gas cleaning, and biodiesel production catalyst.178,179,182–184 Packialakshmi et al.185 showed that coconut shell biochar can be used as an adsorbent for electroplating wastewater, effectively reducing the zinc and potassium concentrations. Moreover, biochar can function as an electrode material for energy production and energy storage devices due to its large surface area and conductivity.186 For instance, Zhang et al.187 used three biomass feedstocks (dairy manure, sewage sludge, and wood chips) to fabricate carbon electrodes. Their study showed that biochar can be converted into carbon electrodes, with their oxygen reduction reaction activity influenced by the feedstock type and pyrolysis temperature. Another recent study by Nambyaruveettil et al.188 developed and optimized a novel Ni-zeolite-biochar hybrid catalyst for green hydrogenation using mordenite zeolite and date pit biochar, presenting an innovative and eco-friendly flexible catalytic support. To utilize biochar in a broad range of applications, its surface area and functional groups need to be upgraded by applying several techniques such as physical and chemical activation, gas activation, and steam activation.180,181,189,190

The gas mixture produced during pyrolysis primarily consists of CO2, CO, H2, CH4, C2H4, and C2H6, along with smaller amounts of NH3, C3H8, sulfur oxides, nitrogen oxides, and low-carbon alcohols. Its typical energy content ranges from 10 to 20 MJ m−3. However, for practical applications, impurities such as aerosols, water vapor, NH3, HCN, and H2S must first be removed.191 The pyrolytic gas composition is influenced by factors such as temperature, residence time, and properties of the biomass feedstock used in the pyrolysis process.192 Light hydrocarbons, particularly CH4, originate from the decomposition of weak methoxyl (–O–CH3) and methylene (–CH2–) groups and the secondary breakdown of oxygenated compounds, while H2 is formed via secondary decomposition and reforming of aromatic C[double bond, length as m-dash]C and C–H bonds at elevated temperature.193 However, the thermal degradation of the carbonyl and carboxyl groups in the three lignocellulosic components is thought to produce CO2 and CO.194 Pyrolytic gas can be utilized for heat and power generation in engines, as a self-feeding source in continuous pyrolysis reactors, or as a carrier gas in fluidized bed systems. However, its short combustion time, difficult storage, and thermal instability limit its efficient and long-term use.191,194 Yu et al.195 highlighted that syngas purification and CO2 capture can significantly enhance the efficiency and value of pyrolysis gas utilization, particularly if the captured CO2 is further synthesized. Another study by Straka et al.196 found that pyrolysis gas from apricot stones and walnut shells can be converted into H2-rich gas using a Pd/C catalyst, whereas CH4-rich gas can be produced using an Ru/Al2O3 catalyst in a sealed pressurized reactor, emphasizing that pressure-catalyzed biomass pyrolysis can efficiently produce energy-rich gas. A recent study by Brands et al.197 explored an innovative approach to convert pyrolysis gas into biomethane through a water-gas shift reaction and microbial methanogenesis.

7.1. Bio-oil upgrading

Bio-oil upgrading is essential for its utilization as a fossil fuel alternative.13,198,199 The major bio-oil catalytic upgrading reaction pathways include dehydration, decarboxylation, and decarbonylation, as depicted in Fig. 16. Additionally, other reactions contribute to bio-oil upgrading such as aromatization, ketonization, cracking, and aldol condensation. Dehydration reaction takes place on the catalyst active sites, producing dehydrated products and water. Decarboxylation reaction expels oxygen from fatty acids in the form of CO2, and decarbonylation removes the carbonyl group from ketones and aldehydes, producing CO. Ketonization involves the reaction of two carboxylic acids to produce ketone, CO2, and water by coupling a C–C bond and removing one carboxyl group.200,201 The aromatization reaction for alkenes and low molecular weight oxygenated compounds such as furans produces aromatic hydrocarbon, CO2, CO, and water. This process can be interpreted through the hydrocarbon pool mechanism.3,201 Soldatos et al.202 investigated the in situ catalytic upgrading of beechwood lignin pyrolysates. Various aluminosilicate catalysts with different Si/Al ratios were employed, among which ZSM-5 (40) demonstrated a superior performance in producing BTX aromatics, while suppressing further condensation into PAHs. Chaerusan et al.203 investigated the catalytic upgrading of bio-oil from torrefied giant Miscanthus over copper-magnesium (Cu-Mg) bimetal-doped zeolites. The results showed a high yield of aromatic hydrocarbons (63%), predominantly favoring BTX compounds. Investigating the variation in surface electronic and structural properties during a specific process is crucial for developing new catalytic systems. X-ray photoelectron spectroscopy (XPS) is an analytical technique used for the surface species and valence analysis of catalysts; it is based on the photoelectric effect, which occurs when a material is exposed to high-energy photons, causing electrons emission.204 Scanning electron microscopy (SEM) is another analysis employed to explore the morphological characteristics of catalysts. XPS and SEM analyses have been employed in several studies,205 for instance Ismail et al.206 utilized the XPS and SEM techniques to explore the chemical state and morphology of a Co-MoS2 catalyst, as shown in Fig. 17. Similarly, Ali et al.207 investigated the characteristics of a 5%Ni-CeO2 catalyst via the aforementioned analyses.
image file: d5ra05770a-f16.tif
Fig. 16 Major reaction pathways for bio-oil catalytic upgrading.

image file: d5ra05770a-f17.tif
Fig. 17 XPS spectra and SEM images of the Co-MoS2 catalyst, adapted from ref. 206 with permission from Elsevier, Case Studies in Chemical and Environmental Engineering, vol. 9, p. 100734, Copyright 2024.

Hydrodeoxygenation (HDO) is a catalytic upgrading process conducted under high hydrogen pressure (70–90 bar) and temperature (<400 °C), which aims to lower the oxygen level in bio-oil by reacting its oxygenated functional groups with hydrogen, acting as a reducing agent. HDO is considered the most efficient bio-oil upgrading treatment.208,209 ZSM-5 zeolite and sulfided CoMo and NiMo catalysts are commonly employed in hydrotreating bio-oil, where CoMo and NiMo are resistant to the sulfur content in the biomass feedstock.210 ZSM-5 and HZSM-5 zeolite catalysts show superior deoxygenation ability owing to their high acidity and accessible acid sites.211 Catalyst deactivation due to coke formation is the primary challenge in HDO treatment.12 Du et al.212 proposed a novel interfacial-enhanced HDO system for upgrading vanillin bio-oil. Hydrophobic silica was used as a solid emulsifier to construct a stable oil/water emulsion, enabling the interfacial HDO of vanillin with an Ni-based catalyst. This approach improved the conversion and increased the deoxygenated selectivity from 43% to 99% at 150 °C, demonstrating a superior performance compared to direct HDO. Xie et al.213 synthesized a series of low-loading Ni/C catalysts (1–7%) for the HDO of lignin bio-oil and phenolic compounds. Using 5% Ni/C at 200 °C for 2 h, guaiacol was completely converted to cyclohexanol, which can serve both as a source for chemical products and as a fossil fuel additive. Several studies related to bio-oil upgrading by HDO are summarized in Table 13. Besides HDO, there are several advanced bio-oil upgrading processes, as follows:

Table 13 Studies on HDO treatment
Biomass Reactor Catalyst Main findings References
Guaiacol as a model compound Autoclave reactor Ni-CoOx/Hβ-u catalysts Under a hydrogen pressure of 1.0 MPa, guaiacol was completely converted to cyclohexane within 40 min at 180 °C 224
Eucalyptus-derived lignin oils Autoclave Mo/NiFe71-V -The catalytic performance was influenced by Mo-Ni intermetallic interactions via charge transfer, with Mo/NiFe71-V showing excellent activity across a wide range of substrates 225
-HDO of guaiacol at 240 °C achieved almost complete conversion, 99.5% cyclohexane yield, and 92.1% yield after six cycles
Pine sawdust Fluidized-bed reactor Fe-Co/SiO2 22.44% was the highest HC content in the upgraded bio-oil 226
Wood Autoclave Rh, Pt, Pd on a zirconia support 37–47 wt% yield of biofuel 227
Loblolly pine Auger pyrolysis reactor Nickel/silica–alumina catalyst HDO for the obtained bio-oil produced HC yield of 20.4 wt% 228
1,6-Anhydro-beta-D glucose Tubular reactor Ni-CeO2 Jet fuel fractions account for the highest share (47.0–68.0%), while gasoline components range from ∼9.0–22.9% 229
Acetone and phenol as model compounds Autoclave Raney Ni catalyst The content of ketones and aldehydes reduced from 30.36% to 9.32% 230
Pine wood Packed bed reactor Ru/C and Pt/C catalysts 45% of the water-soluble carbon in the bio-oil was converted to gasoline blend stocks and C2 to C6 diols 231
Salicornia seeds TGA 5 %Ni-CeO2 Jet fuel compounds yield ranged from 42–45% within temperature range 400–500 °C 207
Date palm pits 2-Stage pyrolysis reactor Co-MoS2 Biomass was converted into non-oxygenated aromatics at temperature 500 °C 206
Date palm pits Multi-stage catalytic pyrolysis reactor Ni-doped-H-beta catalyst BXT compounds (mainly toluene) peaked at 70% between 300–400 °C 232


• Solvent extraction by adding solvents such as methanol, ethanol, and acetone to reduce viscosity of bio-oil and enhance the heating values.214 Machado et al.215 assessed bio-oil deacidification via liquid–liquid extraction with aqueous methanol, showing that higher water and acid contents reduce the efficiency, while higher temperature enhances it. Another study by Zhao et al.216 evaluated the impact of different extraction solvents (acetone, ethanol, dichloromethane and ethyl acetate) on the recovery of bio-oil from rubberwood sawdust liquefaction in water, ethanol or water–ethanol mixed solvent. Among them, acetone achieved the highest oil yield with the water–ethanol solvent, while ethanol was the least effective.

• Emulsification by blending bio-oil with refinery distillates, with the aid of surfactants, to enhance the ignition characteristics.217,218 Bhat et al.219 emulsified bio-crude from the hydrothermal liquefaction of waste wheat flour and canola meal with light cycle oil (LCO) using octanol as an emulsifier. The resulting emulsion showed improved physicochemical properties, including acid number, density, viscosity, heating value, pour point, and elemental composition, compared to the raw bio-crude. Chong et al.220 produced bio-oil from palm kernel shells and developed a stable bio-oil/diesel emulsion via ultrasonic emulsification. Among the tested solvents (2-octanol, 2-heptanol, and 2-octanone), which were selected based on the computer-aided molecular design (CAMD) method, 2-octanol was identified as the most effective for stable emulsions. The optimal diesel:surfactant:bio-oil ratio was 80[thin space (1/6-em)]:[thin space (1/6-em)]15[thin space (1/6-em)]:[thin space (1/6-em)]5, preventing phase separation even at high temperatures. However, although promising, further evaluation of the fuel properties is needed for diesel engine applications.

• Supercritical fluids (SCFs) exhibit solvent properties that allow them to dissolve compounds that are not soluble in either the liquid or gas phase of the solvent. SCFs improve the calorific value and lower the bio-oil viscosity.127,214 Li et al.221 developed an eco-friendly supercritical CO2-ethanolysis system to convert Caragana korshinskii into valuable bio-oil. The process showed strong synergistic degradation, yielding 38.2 wt% bio-oil with a high heating value.

• Plasma processing generates reactive species, such as electrons, using high voltages (10–30 kV) and pulse frequencies (1–10 kHz), promoting the hydrodeoxygenation of bio-oil and improving the fuel quality. When combined with bifunctional catalysts or processes such as biomass-plastic co-pyrolysis, it enhances the H/C ratio. However, challenges including high costs, technical complexity, and catalyst deactivation from coke require further research.222 Fan et al.223 studied the catalytic pyrolysis of rapeseed shells using HZSM-5 coupled with plasma discharge, producing bio-oil with a high purity of monocyclic aromatic hydrocarbons and no detectable polycyclic aromatic hydrocarbons, and with its C8–C12 carbon range, the refined bio-oil is suitable as a gasoline or diesel additive.

8. Challenges and future prospects

Biomass valorization into fossil fuel substitutes through thermochemical conversion faces several challenges and presents prospects for further study. The high oxygen content in the produced bio-oil is the main challenge in biomass conversion, as it restricts its direct application as a fossil fuel substitute and leads to inferior quality compared to petroleum-based fuels.1,233 As discussed earlier, HDO is the most effective upgrading method as it increases the calorific value and enhances the fuel quality by reducing the oxygen content through reactions including cracking, decarbonylation, decarboxylation, and hydrogenation, with catalysts playing a key role.234 However, although noble metals such as Pt, Pd, Ru, and Rh show high activity in HDO, their high cost and scarcity limit their industrial application. Transition metals such as Ni, Fe, and Co, either alone or in combination and supported on suitable materials, can achieve comparable HDO performances. Among them, Ni is the most widely used due to its low cost, strong hydrogenation ability, and efficient C–O bond cleavage.235 A recent study by Wu et al.236 reported that Ni-V bimetallic zeolite catalysts effectively upgraded pine nut shell pyrolysis vapors via in situ HDO. The treatment improved the pH and lowered the heating value of the bio-oil, reduced its moisture content and density, lowered its O/C ratio, and efficiently reduced its content of oxygenated compounds. Catalyst deactivation remains a major challenge, primarily driven by carbon deposition, sintering, and poisoning of the active sites. These processes diminish the efficiency of HDO catalysts by lowering their available surface area and modifying their chemical properties.234 Designing nanocatalysts with a core–shell structure is a promising approach to enhance the catalyst stability, as it helps prevent nanoparticle migration and sintering.237 Deng et al.238 showed that an Ru@SiO2-ZrO2 core–shell catalyst outperformed conventional Ru/SiO2-ZrO2, maintaining its activity over five cycles without notable deactivation. This catalyst displayed an excellent HDO performance for phenolic compounds and lignin pyrolysis oil, increasing its hydrocarbon content from 12.7% to 96.9%, and thus highlighting its strong potential for industrial application. Catalysts often suffer deactivation from carbon deposition. Thus, common approaches such as thermal oxidation, steam treatment, and gasification are primarily employed to remove coke, poisons, or surface deposits, which diminish the catalyst activity. However, although widely used, these methods typically require significant energy and can potentially damage the structural integrity of the catalyst. Also, the conventional regeneration methods often fail to fully recover the catalyst activity, prompting the development of emerging techniques, such as plasma treatment and microwave-assisted regeneration to overcome the severe deactivation, structural damage, and loss of selectivity in catalysts.239,240 Non-thermal plasma (NTP) has emerged as an efficient alternative for catalyst regeneration, enabling non-thermal molecular activation and the generation of highly reactive species under ambient pressure and near-room temperature conditions.241 During NTP, the excited radicals oxidize and gasify coke on the catalyst at low temperatures, forming CO and CO2 without overheating the catalyst.240 Li et al.242 showed that direct NTP regeneration effectively restored the activity of deactivated HZSM-5 for upgrading rape straw pyrolysis vapors by removing coke and generating active sites, while also increasing the surface area, pore volume, and acidity of the catalyst.

Mechanical pretreatment is essential for most biomass conversion processes because biomass is typically a heterogeneous solid with varying sizes, shapes, and compositions, which can reduce the conversion efficiency and complicate handling and transport. Processes such as crushing, grinding, drying, and densification help produce a more uniform feedstock, improving the material handling and overall conversion performance.90,243 The particle size of biomass feedstock significantly impacts bio-oil production, influencing both the yield and quality. Fast pyrolysis favors smaller particles, which enable efficient and uniform heat transfer, whereas larger particles hinder heat penetration, causing lower internal temperatures, and consequently reduced liquid yields, and thus require additional mechanical energy for feedstock physical pretreatment.244–246 Biomass feedstocks for pyrolysis often contain over 40% moisture, necessitating pre-treatment such as atmospheric drying, solar drying, or evaporation. Although solar drying is economical, it is time-consuming. Also, the high water content in biomass increases the energy consumption for evaporation, thereby limiting the process efficiency and economic viability.247

Techno-economic analysis (TEA) is vital for evaluating the feasibility and sustainability of bio-oil production. It quantifies costs, profitability, and long-term viability. Given that fossil fuels remain cheaper, bio-oil must demonstrate cost competitiveness and economic sustainability to scale successfully. Its economic feasibility can be enhanced through flexible feedstock use, co-product valorization, and efficient upgrading methods that lower costs and improve the fuel quality.69,148 Additionally, it is crucial to optimize a system that can handle the variation in biomass feedstock materials. The inconsistency in biomass can be mitigated by developing standards for the feedstocks that can be defined based on parameters such as moisture content and LHV.20,39

The use of AI in biorefineries is limited by poor data quality and insufficient data quantity. Hybrid models, which combine data-driven and mechanistic approaches, reduce data needs, while advanced sensors improve data quality and availability. Soft sensors further address these gaps by estimating unmeasured variables in real time, enabling timely interventions and ensuring accurate outcomes.156 Current ML studies in thermochemical conversion mainly predict single bioenergy products, overlooking competitive reactions among biochar, bio-oil, and biogas. Thus, future research should address these interactions to optimize the yields and clarify the mechanisms.153

9. Conclusions

This study reviews the valorization of biomass materials into bioenergy through thermochemical processes (pyrolysis, gasification, liquefaction, and combustion). It can be inferred that biomass is a promising source for bio-oil, syngas, biochar, and heat, which can contribute into minimizing the reliance on fossil fuel sources and satisfy the global demand for renewable energy. Pyrolysis and hydrothermal liquefaction have demonstrated strong potential for converting lignocellulosic biomass into liquid bio-oils, which can be further upgraded into biofuels. The gasification process can be used to produce hydrogen, an emerging renewable energy source with applications such as fuel cell electric vehicles. Thermochemical techniques encounter several challenges that restrict their direct application as a sustainable and commercial energy source. The complexity of biomass feedstock complicates the optimization of systems able to treat the inconsistent feed. Moreover, the formation of oxygenated compounds, which reduce the fuel calorific value, calls the upgradation of bio-oil. HDO is the most effective upgrading technique. Nevertheless, the short lifetime of the catalyst due to its deactivation is the main challenge faced. Future research should prioritize emerging techniques such as non-thermal plasma and microwave-assisted regeneration to mitigate severe catalyst deactivation and associated structural damage. Additionally, developing nanocatalysts with a core–shell structure offers a promising strategy to improve the catalyst stability. Blending bio-oil with petroleum-based fuels through emulsification supports a more economical path for biofuel production. Further research should emphasize the integration of solar energy and biomass conversion technologies to tackle climate change and promote a shift toward cleaner energy sources. Although ML has shown success in bioenergy production, its use is still emerging and largely limited to lab-scale systems. Further research is needed to apply AI in the thermochemical valorization of biomass for its commercial and sustainable conversion.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results, software or code have been included, and no new data were generated or analysed as part of this review.

Acknowledgements

This study was supported by a grant from the United Arab Emirates University, grant ID: 12N166.

References

  1. N. C. Joshi, et al., A concise review on waste biomass valorization through thermochemical conversion, Curr. Res. Microb. Sci., 2024, 6, 100237,  DOI:10.1016/J.CRMICR.2024.100237 .
  2. A. Garba, Biomass Conversion Technologies for Bioenergy Generation: An Introduction, Biotechnological Applications of Biomass, 2020,  DOI:10.5772/INTECHOPEN.93669 .
  3. S. Wang, G. Dai, H. Yang, and Z. Luo, Lignocellulosic Biomass Pyrolysis Mechanism: A State-Of-The-Art Review, Elsevier Ltd, 2017,  DOI:10.1016/j.pecs.2017.05.004 .
  4. Biofuels – Renewables 2021 – Analysis - IEA, Accessed: Feb. 11, 2025. [Online]. Available: https://www.iea.org/reports/renewables-2021/biofuels?mode=transport%26region=World%26publication=2021%26flow=Consumption%26product=Ethanol Search PubMed.
  5. S. Jha, S. Nanda, B. Acharya and A. K. Dalai, A Review of Thermochemical Conversion of Waste Biomass to Biofuels, Energies, 2022, 15, 6352,  DOI:10.3390/EN15176352 .
  6. J. Cai, et al., Research on the application of catalytic materials in biomass pyrolysis, J. Anal. Appl. Pyrolysis, 2024, 177, 106321,  DOI:10.1016/J.JAAP.2023.106321 .
  7. G. Koçar and N. Civaş, An overview of biofuels from energy crops: Current status and future prospects, Renew. Sustain. Energy Rev., 2013, 28, 900–916,  DOI:10.1016/J.RSER.2013.08.022 .
  8. A. Demirbaş, Biomass resource facilities and biomass conversion processing for fuels and chemicals, Energy Convers. Manag., 2001, 42(11), 1357–1378,  DOI:10.1016/S0196-8904(00)00137-0 .
  9. M. Madadi, et al., Transformative biorefinery model for biomass valorization into biofuel and renewable platform chemicals, J. Energy Chem., 2025, 110, 109–123,  DOI:10.1016/J.JECHEM.2025.06.016 .
  10. Y. Saleh, L. Ali, A. Alam, M. Kuttiyathil and M. Altarawneh, Pyrolysis characteristics, thermo-kinetic parameters, and evolved gas analysis for waste biomass derived from food processing industries, Green Technol. Sustain., 2025, 100266,  DOI:10.1016/J.GRETS.2025.100266 .
  11. N. Canabarro, J. F. Soares, C. G. Anchieta, C. S. Kelling, and M. A. Mazutti, Thermochemical processes for biofuels production from biomass, 2013, [Online]. Available: http://www.sustainablechemicalprocesses.com/content/1/1/22 Search PubMed.
  12. X. Hu and M. Gholizadeh, Progress of the applications of bio-oil, Renew. Sustain. Energy Rev., 2020, 134, 110124,  DOI:10.1016/J.RSER.2020.110124 .
  13. S. Xiu and A. Shahbazi, Bio-oil production and upgrading research: A review, Renew. Sustain. Energy Rev., 2012, 16(7), 4406–4414,  DOI:10.1016/J.RSER.2012.04.028 .
  14. H. Machado, A. F. Cristino, S. Orišková and R. Galhano dos Santos, Bio-Oil: The Next-Generation Source of Chemicals, Reactions, 2022, 3(1), 118–137,  DOI:10.3390/REACTIONS3010009 .
  15. M. khan, et al., Applications of machine learning in thermochemical conversion of biomass-A review, Fuel, 2023, 332, 126055,  DOI:10.1016/J.FUEL.2022.126055 .
  16. N. Dahmen, I. Lewandowski, S. Zibek and A. Weidtmann, Integrated lignocellulosic value chains in a growing bioeconomy: Status quo and perspectives, GCB Bioenergy, 2019, 11(1), 107–117,  DOI:10.1111/GCBB.12586 .
  17. C. C. Seah, et al., Co-pyrolysis of biomass and plastic: Circularity of wastes and comprehensive review of synergistic mechanism, Results Eng., 2023, 17, 100989,  DOI:10.1016/J.RINENG.2023.100989 .
  18. N. Tripathi, C. D. Hills, R. S. Singh and C. J. Atkinson, Biomass waste utilisation in low-carbon products: harnessing a major potential resource, npj Clim. Atmos. Sci., 2019, 2(1), 1–10,  DOI:10.1038/s41612-019-0093-5 .
  19. B. Koul, M. Yakoob and M. P. Shah, Agricultural waste management strategies for environmental sustainability, Environ. Res., 2022, 206, 112285,  DOI:10.1016/J.ENVRES.2021.112285 .
  20. P. P. Paudel, S. Kafle, S. Park, S. J. Kim, L. Cho and D. H. Kim, Advancements in sustainable thermochemical conversion of agricultural crop residues: A systematic review of technical progress, applications, perspectives, and challenges, Renew. Sustain. Energy Rev., 2024, 202, 114723,  DOI:10.1016/J.RSER.2024.114723 .
  21. K. M. Akkoli, P. B. Gangavati, M. R. Ingalagi and R. K. Chitgopkar, Assessment and Characterization of Agricultural Residues, Mater. Today Proc., 2018, 5(9), 17548–17552,  DOI:10.1016/J.MATPR.2018.06.071 .
  22. M. Raza, L. Ali, M. Altarawneh and B. Abu-Jdayil, Evaluation of thermal pyrolysis characteristics and kinetic parameters of cellulose extracted from date waste using natural deep eutectic solvent, Case Stud. Chem. Environ. Eng., 2024, 10, 100796,  DOI:10.1016/J.CSCEE.2024.100796 .
  23. A. Baiano, Recovery of Biomolecules from Food Wastes — A Review, Molecules, 2014, 19(9), 14821–14842,  DOI:10.3390/MOLECULES190914821 .
  24. S. Kumar and J. Konwar, Current progress in valorization of food processing waste and by-products for pectin extraction, Int. J. Biol. Macromol., 2023, 239, 124332 Search PubMed .
  25. K. Selva Ganesh, A. Sridhar and S. Vishali, Utilization of fruit and vegetable waste to produce value-added products:Conventional utilization and emerging opportunities-A review, Chemosphere, 2022, 287, 132221 Search PubMed .
  26. E. Capanoglu, E. Nemli and F. Tomas-Barberan, Novel Approaches in the Valorization of Agricultural Wastes and Their Applications, J. Agric. Food Chem., 2022, 70(23), 6787–6804,  DOI:10.1021/ACS.JAFC.1C07104 .
  27. A. Jezierska-Thöle, R. Rudnicki and M. Kluba, Development of energy crops cultivation for biomass production in Poland, Renew. Sustain. Energy Rev., 2016, 62, 534–545,  DOI:10.1016/J.RSER.2016.05.024 .
  28. L. López-Bellido, J. Wery and R. J. López-Bellido, Energy crops: Prospects in the context of sustainable agriculture, Eur. J. Agron., 2014, 60, 1–12,  DOI:10.1016/J.EJA.2014.07.001 .
  29. M. Parikka, Global biomass fuel resources, Biomass Bioenergy, 2004, 27(6), 613–620,  DOI:10.1016/J.BIOMBIOE.2003.07.005 .
  30. Y. Li, L. W. Zhou and R. Z. Wang, Urban biomass and methods of estimating municipal biomass resources, Renew. Sustain. Energy Rev., 2017, 80, 1017–1030,  DOI:10.1016/J.RSER.2017.05.214 .
  31. Y. Tan, et al., Integrated biorefinery routes to transform furfural waste into 2G biofuels and PFOA-adsorbing biochar, Green Chem., 2025, 27(13), 3573–3589,  10.1039/D5GC00199D .
  32. Y. Du, T. Ju, Y. Meng, T. Lan, S. Han and J. Jiang, A review on municipal solid waste pyrolysis of different composition for gas production, Fuel Process. Technol., 2021, 224, 107026,  DOI:10.1016/J.FUPROC.2021.107026 .
  33. Y. K N, et al., Lignocellulosic biomass-based pyrolysis: A comprehensive review, Chemosphere, 2022, 286, 131824,  DOI:10.1016/j.chemosphere.2021.131824 .
  34. R. S. Riseh, M. G. Vazvani, M. Hassanisaadi and V. K. Thakur, Agricultural wastes: A practical and potential source for the isolation and preparation of cellulose and application in agriculture and different industries, Ind. Crops Prod., 2024, 208, 117904,  DOI:10.1016/J.INDCROP.2023.117904 .
  35. J. Rao, Z. Lv, G. Chen and F. Peng, Hemicellulose: Structure, chemical modification, and application, Prog. Polym. Sci., 2023, 140, 101675,  DOI:10.1016/J.PROGPOLYMSCI.2023.101675 .
  36. S. D. Stefanidis, K. G. Kalogiannis, E. F. Iliopoulou, C. M. Michailof, P. A. Pilavachi and A. A. Lappas, A study of lignocellulosic biomass pyrolysis via the pyrolysis of cellulose, hemicellulose and lignin, J. Anal. Appl. Pyrolysis, 2014, 105, 143–150,  DOI:10.1016/J.JAAP.2013.10.013 .
  37. L. Zhang, C. Charles) Xu and P. Champagne, Overview of recent advances in thermo-chemical conversion of biomass, Energy Convers. Manag., 2010, 51(5), 969–982,  DOI:10.1016/J.ENCONMAN.2009.11.038 .
  38. Biomass Conversion Technology: Deep Dive | Nexus PMG, Accessed: Feb. 15, 2025. [Online]. Available: https://www.nexuspmg.com/insights/biomass-conversion-technology-overview Search PubMed.
  39. V. Dhyani and T. Bhaskar, A comprehensive review on the pyrolysis of lignocellulosic biomass, Renew. Energy, 2018, 129, 695–716,  DOI:10.1016/j.renene.2017.04.035 .
  40. J. Liu, et al., Biomass pyrolysis mechanism for carbon-based high-value products, Proc. Combust. Inst., 2023, 39(3), 3157–3181,  DOI:10.1016/j.proci.2022.09.063 .
  41. D. Chen, et al., Insight into biomass pyrolysis mechanism based on cellulose, hemicellulose, and lignin: Evolution of volatiles and kinetics, elucidation of reaction pathways, and characterization of gas, biochar and bio-oil, Combust. Flame, 2022, 242, 112142,  DOI:10.1016/j.combustflame.2022.112142 .
  42. D. Neves, H. Thunman, A. Matos, L. Tarelho and A. Gómez-Barea, Characterization and prediction of biomass pyrolysis products, Prog. Energy Combust. Sci., 2011, 37(5), 611–630,  DOI:10.1016/J.PECS.2011.01.001 .
  43. P. Wang, S. Zhan, H. Yu, X. Xue and N. Hong, The effects of temperature and catalysts on the pyrolysis of industrial wastes (herb residue), Bioresour. Technol., 2010, 101(9), 3236–3241,  DOI:10.1016/J.BIORTECH.2009.12.082 .
  44. D. Chen, D. Liu, H. Zhang, Y. Chen and Q. Li, Bamboo pyrolysis using TG–FTIR and a lab-scale reactor: Analysis of pyrolysis behavior, product properties, and carbon and energy yields, Fuel, 2015, 148, 79–86,  DOI:10.1016/J.FUEL.2015.01.092 .
  45. U. Morali and S. Şensöz, Pyrolysis of hornbeam shell (Carpinus betulus L.) in a fixed bed reactor: Characterization of bio-oil and bio-char, Fuel, 2015, 150, 672–678,  DOI:10.1016/J.FUEL.2015.02.095 .
  46. G. Kabir, A. T. Mohd Din and B. H. Hameed, Pyrolysis of oil palm mesocarp fiber and palm frond in a slow-heating fixed-bed reactor: A comparative study, Bioresour. Technol., 2017, 241, 563–572,  DOI:10.1016/J.BIORTECH.2017.05.180 .
  47. A. M. Parvej, M. A. Rahman and K. M. A. Reza, The combined effect of solar assisted torrefaction and pyrolysis on the production of valuable chemicals obtained from water hyacinth biomass, Clean. Waste Syst., 2022, 3, 100027,  DOI:10.1016/J.CLWAS.2022.100027 .
  48. H. Elshareef, et al., Investigation of Bio-Oil and Biochar Derived from Cotton Stalk Pyrolysis: Effect of Different Reaction Conditions, Resources, 2025, 14, 75,  DOI:10.3390/RESOURCES14050075 .
  49. M. Kumar, S. N. Upadhyay and P. K. Mishra, Pyrolysis of Sugarcane (Saccharum officinarum L.) Leaves and Characterization of Products, ACS Omega, 2022, 7(32), 28052–28064,  DOI:10.1021/ACSOMEGA.2C02076/ASSET/IMAGES/LARGE/AO2C02076_0009.JPEG .
  50. D. Chen, D. Liu, H. Zhang, Y. Chen and Q. Li, Bamboo pyrolysis using TG–FTIR and a lab-scale reactor: Analysis of pyrolysis behavior, product properties, and carbon and energy yields, Fuel, 2015, 148, 79–86,  DOI:10.1016/J.FUEL.2015.01.092 .
  51. K. Akubo, M. A. Nahil and P. T. Williams, Pyrolysis-catalytic steam reforming of agricultural biomass wastes and biomass components for production of hydrogen/syngas, J. Energy Inst., 2019, 92(6), 1987–1996,  DOI:10.1016/J.JOEI.2018.10.013 .
  52. W. C. Wang and A. C. Lee, The study of producing ‘drop-in’ fuels from agricultural waste through fast pyrolysis and catalytic hydro-processing, Renew. Energy, 2019, 133, 1–10,  DOI:10.1016/J.RENENE.2018.10.022 .
  53. Y. K. S. S. Rao, et al., Investigation on Forestry Wood Wastes: Pyrolysis and Thermal Characteristics of Ficus religiosa for Energy Recovery System, Adv. Mater. Sci. Eng., 2022, 2022(1), 3314606,  DOI:10.1155/2022/3314606 .
  54. M. A. Lopez-Velazquez, V. Santes, J. Balmaseda and E. Torres-Garcia, Pyrolysis of orange waste: A thermo-kinetic study, J. Anal. Appl. Pyrolysis, 2013, 99, 170–177,  DOI:10.1016/j.jaap.2012.09.016 .
  55. H. Haykiri-Acma, S. Yaman and S. Kucukbayrak, Effect of heating rate on the pyrolysis yields of rapeseed, Renew. Energy, 2006, 31(6), 803–810,  DOI:10.1016/J.RENENE.2005.03.013 .
  56. R. S. A. Nounou, I. A. Shaban, L. Ali, M. Altarawneh and M. Elgendi, Kinetic and thermodynamic analysis of the PET-date seed mixture co-pyrolysis using Coats-Redfern method, Renew. Energy, 2024, 237, 121524,  DOI:10.1016/J.RENENE.2024.121524 .
  57. Z. Wang, K. G. Burra, T. Lei and A. K. Gupta, Co-pyrolysis of waste plastic and solid biomass for synergistic production of biofuels and chemicals-A review, Prog. Energy Combust. Sci., 2021, 84, 100899,  DOI:10.1016/J.PECS.2020.100899 .
  58. S. Y. Oh and Y. D. Seo, Polymer/biomass-derived biochar for use as a sorbent and electron transfer mediator in environmental applications, Bioresour. Technol., 2016, 218, 77–83,  DOI:10.1016/J.BIORTECH.2016.06.073 .
  59. V. Kudelytė, J. Eimontas, R. Paulauskas and N. Striūgas, Co-Pyrolysis of Plastic Waste and Lignin: A Pathway for Enhanced Hydrocarbon Recovery, Energies, 2025, 18, 275,  DOI:10.3390/EN18020275 .
  60. E. Önal, B. B. Uzun and A. E. Pütün, Bio-oil production via co-pyrolysis of almond shell as biomass and high density polyethylene, Energy Convers. Manag., 2014, 78, 704–710,  DOI:10.1016/J.ENCONMAN.2013.11.022 .
  61. J. D. Martínez, et al., Co-pyrolysis of biomass with waste tyres: Upgrading of liquid bio-fuel, Fuel Process. Technol., 2014, 119, 263–271,  DOI:10.1016/J.FUPROC.2013.11.015 .
  62. L. M. Terry, et al., Co-pyrolysis of oil palm trunk and polypropylene: Pyrolysis oil composition and formation mechanism, S. Afr. J. Chem. Eng., 2023, 43, 348–358,  DOI:10.1016/J.SAJCE.2022.12.001 .
  63. R. Dong, et al., Co-pyrolysis of vineyards biomass waste and plastic waste: Thermal behavior, pyrolytic characteristic, kinetics, and thermodynamics analysis, J. Anal. Appl. Pyrolysis, 2024, 179, 106506,  DOI:10.1016/j.jaap.2024.106506 .
  64. M. Brebu and I. Spiridon, Co-pyrolysis of LignoBoost® lignin with synthetic polymers, Polym. Degrad. Stab., 2012, 97(11), 2104–2109,  DOI:10.1016/J.POLYMDEGRADSTAB.2012.08.024 .
  65. A. O. Oyedun, C. Z. Tee, S. Hanson and C. W. Hui, Thermogravimetric analysis of the pyrolysis characteristics and kinetics of plastics and biomass blends, Fuel Process. Technol., 2014, 128, 471–481,  DOI:10.1016/J.FUPROC.2014.08.010 .
  66. Ö. Çepelioǧullar and A. E. Pütün, Thermal and kinetic behaviors of biomass and plastic wastes in co-pyrolysis, Energy Convers. Manag., 2013, 75, 263–270,  DOI:10.1016/j.enconman.2013.06.036 .
  67. M. M. Hasan, M. G. Rasul, M. M. K. Khan, N. Ashwath and M. I. Jahirul, Energy recovery from municipal solid waste using pyrolysis technology: A review on current status and developments, Renew. Sustain. Energy Rev., 2021, 145, 111073,  DOI:10.1016/J.RSER.2021.111073 .
  68. P. Talwar, M. A. Agudelo and S. Nanda, Pyrolysis Process, Reactors, Products, and Applications: A Review, Energies, 2025, 18, 2979,  DOI:10.3390/EN18112979 .
  69. W. Jerzak, E. Acha and B. Li, Comprehensive Review of Biomass Pyrolysis: Conventional and Advanced Technologies, Reactor Designs, Product Compositions and Yields, and Techno-Economic Analysis, Energies, 2024, 17, 5082,  DOI:10.3390/EN17205082 .
  70. J. P. Bok, Y. S. Choi, S. K. Choi and Y. W. Jeong, Fast pyrolysis of Douglas fir by using tilted-slide reactor and characteristics of biocrude-oil fractions, Renew. Energy, 2014, 65, 7–13,  DOI:10.1016/J.RENENE.2013.06.035 .
  71. F. Campuzano, R. C. Brown and J. D. Martínez, Auger reactors for pyrolysis of biomass and wastes, Renew. Sustain. Energy Rev., 2019, 102, 372–409,  DOI:10.1016/J.RSER.2018.12.014 .
  72. Temperature Evolution and Heating Rates of Biomass Undergoing Ablative Pyrolysis | Engineering, Technology & Applied Science Research. Accessed: Jan. 19, 2025. [Online]. Available: https://etasr.com/index.php/ETASR/article/view/5621/3031 Search PubMed.
  73. H. Cong, et al., Comprehensive analysis of industrial-scale heating plants based on different biomass slow pyrolysis technologies: Product property, energy balance, and ecological impact, Clean Eng. Technol., 2022, 6, 100391,  DOI:10.1016/J.CLET.2021.100391 .
  74. X. Hu and M. Gholizadeh, 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,  DOI:10.1016/J.JECHEM.2019.01.024 .
  75. M. Gholizadeh, C. Castro, S. M. Fabrega and F. Clarens, A review on thermochemical based biorefinery catalyst development progress, Sustain. Energy Fuels, 2023, 7(19), 4758–4804,  10.1039/D3SE00496A .
  76. F.-X. Collard, S. Wijeyekoon, and P. Bennett, Commercial Status of Direct Thermochemical Liquefaction Technologies IEA Bioenergy: Task 34 June 2023 Xxxx: Xx IEA Bioenergy: Task XX Month Year Xxxx: Xx Commercial Status of Direct Thermochemical Liquefaction Technologies Title of publication Subtitle of publication, 2023.
  77. Commercial status of direct thermochemical liquefaction technologies – Bioenergy, Accessed: Sep. 19, 2025. [Online]. Available: https://www.ieabioenergy.com/blog/publications/commercial-status-of-direct-thermochemical-liquefaction-technologies/?utm_source=chatgpt.com Search PubMed.
  78. A. Zabaniotou, Agro-residues implication in decentralized CHP production through a thermochemical conversion system with SOFC, Sustain. Energy Technol. Assessments, 2014, 6, 34–50,  DOI:10.1016/J.SETA.2014.01.004 .
  79. A. Molino, S. Chianese and D. Musmarra, Biomass gasification technology: The state of the art overview, J. Energy Chem., 2016, 25(1), 10–25,  DOI:10.1016/J.JECHEM.2015.11.005 .
  80. Z. Y. Kong, T. J. N. Ang, H. C. Ong, T. Shi, C. I. Yeo and A. Yang, Green hydrogen production from oil palm wastes: A techno-economic and environmental perspective, Energy Convers. Manag., 2025, 341, 120026,  DOI:10.1016/J.ENCONMAN.2025.120026 .
  81. Ö. Tezer, N. Karabağ, A. Öngen, C. Ö. Çolpan and A. Ayol, Biomass gasification for sustainable energy production: A review, Int. J. Hydrogen Energy, 2022, 47(34), 15419–15433,  DOI:10.1016/J.IJHYDENE.2022.02.158 .
  82. D. Baruah and D. C. Baruah, Modeling of biomass gasification: A review, Renew. Sustain. Energy Rev., 2014, 39, 806–815,  DOI:10.1016/J.RSER.2014.07.129 .
  83. V. S. Sikarwar, et al., An overview of advances in biomass gasification, Energy Environ. Sci., 2016, 9(10), 2939–2977,  10.1039/C6EE00935B .
  84. M. Puig-Arnavat, J. C. Bruno and A. Coronas, Review and analysis of biomass gasification models, Renew. Sustain. Energy Rev., 2010, 14(9), 2841–2851,  DOI:10.1016/J.RSER.2010.07.030 .
  85. C. Higman, Gasification, Combustion Engineering Issues for Solid Fuel Systems, 2008, pp. 423–468,  DOI:10.1016/B978-0-12-373611-6.00011-2 .
  86. J. J. Hernández, G. Aranda-Almansa and A. Bula, Gasification of biomass wastes in an entrained flow gasifier: Effect of the particle size and the residence time, Fuel Process. Technol., 2010, 91(6), 681–692,  DOI:10.1016/J.FUPROC.2010.01.018 .
  87. J. S. Tumuluru, B. Ghiasi, N. R. Soelberg and S. Sokhansanj, Biomass Torrefaction Process, Product Properties, Reactor Types, and Moving Bed Reactor Design Concepts, Front. Energy Res., 2021, 9, 728140,  DOI:10.3389/FENRG.2021.728140/BIBTEX .
  88. I. Hassan, O. O. Thomas, F. A. Ezekiel, A. L. Adache and T. S. Mogaji, Impact of Torrefaction Process in Elevating the Fuel Properties of Selected Herbaceous Biomass Solid Waste, J. Eng. Adv., 2024, 64–70,  DOI:10.38032/JEA.2024.03.001 .
  89. D. G. Gizaw, et al., Advances in solid biofuels production through torrefaction: Potential biomass, types of torrefaction and reactors, influencing process parameters and future opportunities – A review, Process Saf. Environ. Prot., 2024, 186, 1307–1319,  DOI:10.1016/J.PSEP.2024.04.070 .
  90. R. Singh, R. Kumar, P. Sarangi, A Kovalev and V Vivekanand, Effect of physical and thermal pretreatment of lignocellulosic biomass on biohydrogen production by thermochemical route: A critical review, Bioresour. Technol., 2023, 369, 128458,  DOI:10.1016/J.BIORTECH.2022.128458 .
  91. B. YAN, et al., Biomass torrefaction coupled gasification technology: A critical review, J. Energy Inst., 2025, 123, 102224,  DOI:10.1016/J.JOEI.2025.102224 .
  92. K. A. Abdulyekeen, W. M. A. W. Daud and M. F. A. Patah, Torrefaction of wood and garden wastes from municipal solid waste to enhanced solid fuel using helical screw rotation-induced fluidised bed reactor: Effect of particle size, helical screw speed and temperature, Energy, 2024, 293, 130759,  DOI:10.1016/J.ENERGY.2024.130759 .
  93. E. A. Silveira, et al., Effect of torrefaction severity on the energy recovery from amazonian wood residues for decentralized energy conversion systems, Biomass Bioenergy, 2025, 193, 107515,  DOI:10.1016/J.BIOMBIOE.2024.107515 .
  94. M. S. Hossain, F. Wasima, M. S. I. K. Shawon, M. Mourshed and B. K. Das, Valorization of food waste into hydrogen energy through supercritical water gasification: Generation potential and techno-econo-environmental feasibility assessment, Renew. Energy, 2024, 235, 121382,  DOI:10.1016/J.RENENE.2024.121382 .
  95. T. Sathish, et al., Waste Coconut oil meal to Hydrogen production through combined steam/water gasification with varying operating parameters and NaCl additions, Int. J. Hydrogen Energy, 2025, 107, 350–358,  DOI:10.1016/J.IJHYDENE.2024.01.323 .
  96. Y. Yue, X. Jin and L. Deng, Experimental Study on Properties of Syngas, Tar, and Biochar Derived from Different Gasification Methods, Appl. Sci., 2023, 13(20), 11490,  DOI:10.3390/APP132011490/S1 .
  97. H. Boujjat, S. Rodat and S. Abanades, Solar-hybrid Thermochemical Gasification of Wood Particles and Solid Recovered Fuel in a Continuously-Fed Prototype Reactor, Energies, 2020, 13, 5217,  DOI:10.3390/EN13195217 .
  98. K. Manatura, J. H. Lu, K. T. Wu and H. Te Hsu, Exergy analysis on torrefied rice husk pellet in fluidized bed gasification, Appl. Therm. Eng., 2017, 111, 1016–1024,  DOI:10.1016/J.APPLTHERMALENG.2016.09.135 .
  99. P. Lv, Z. Yuan, C. Wu, L. Ma, Y. Chen and N. Tsubaki, Bio-syngas production from biomass catalytic gasification, Energy Convers. Manag., 2007, 48(4), 1132–1139,  DOI:10.1016/J.ENCONMAN.2006.10.014 .
  100. M. Asadullah, S. I. Ito, K. Kunimori, M. Yamada and K. Tomishige, Biomass Gasification to Hydrogen and Syngas at Low Temperature: Novel Catalytic System Using Fluidized-Bed Reactor, J. Catal., 2002, 208(2), 255–259,  DOI:10.1006/JCAT.2002.3575 .
  101. A. Ayol, O. Tezer Yurdakos and A. Gurgen, Investigation of municipal sludge gasification potential: Gasification characteristics of dried sludge in a pilot-scale downdraft fixed bed gasifier, Int. J. Hydrogen Energy, 2019, 44(32), 17397–17410,  DOI:10.1016/J.IJHYDENE.2019.01.014 .
  102. A. AlNouss, G. McKay and T. Al-Ansari, Production of syngas via gasification using optimum blends of biomass, J. Clean. Prod., 2020, 242, 118499,  DOI:10.1016/J.JCLEPRO.2019.118499 .
  103. J. Zeng, R. Xiao and J. Yuan, High-quality syngas production from biomass driven by chemical looping on a PY-GA coupled reactor, Energy, 2021, 214, 118846,  DOI:10.1016/J.ENERGY.2020.118846 .
  104. E. Shayan, V. Zare and I. Mirzaee, Hydrogen production from biomass gasification; a theoretical comparison of using different gasification agents, Energy Convers. Manag., 2018, 159, 30–41,  DOI:10.1016/J.ENCONMAN.2017.12.096 .
  105. S. Zhang, S. Zhu, H. Zhang, X. Liu and Y. Xiong, High quality H2-rich syngas production from pyrolysis-gasification of biomass and plastic wastes by Ni–Fe@Nanofibers/Porous carbon catalyst, Int. J. Hydrogen Energy, 2019, 44(48), 26193–26203,  DOI:10.1016/J.IJHYDENE.2019.08.105 .
  106. P. Mondal, G. S. Dang and M. O. Garg, Syngas production through gasification and cleanup for downstream applications — Recent developments, Fuel Process. Technol., 2011, 92(8), 1395–1410,  DOI:10.1016/J.FUPROC.2011.03.021 .
  107. W. Jiang, A. Kumar and S. Adamopoulos, Liquefaction of lignocellulosic materials and its applications in wood adhesives—A review, Ind. Crops Prod., 2018, 124, 325–342,  DOI:10.1016/J.INDCROP.2018.07.053 .
  108. H. J. Huang and X. Z. Yuan, Recent progress in the direct liquefaction of typical biomass, Prog. Energy Combust. Sci., 2015, 49, 59–80,  DOI:10.1016/J.PECS.2015.01.003 .
  109. A. R. K. Gollakota, N. Kishore and S. Gu, A review on hydrothermal liquefaction of biomass, Renew. Sustain. Energy Rev., 2018, 81, 1378–1392,  DOI:10.1016/J.RSER.2017.05.178 .
  110. A. Dimitriadis and S. Bezergianni, Hydrothermal liquefaction of various biomass and waste feedstocks for biocrude production: A state of the art review, Renew. Sustain. Energy Rev., 2017, 68, 113–125,  DOI:10.1016/J.RSER.2016.09.120 .
  111. R. Beauchet, et al., Hydroliquefaction of green wastes to produce fuels, Bioresour. Technol., 2011, 102(10), 6200–6207,  DOI:10.1016/J.BIORTECH.2011.02.043 .
  112. K. M. Isa, T. A. T. Abdullah and U. F. M. Ali, Hydrogen donor solvents in liquefaction of biomass: A review, Renew. Sustain. Energy Rev., 2018, 81, 1259–1268,  DOI:10.1016/J.RSER.2017.04.006 .
  113. C. A. Moreira-Mendoza, et al., Biocrude upgrading with tandem catalysts via hydrothermal liquefaction of biomass feedstocks, Catal. Today, 2026, 461, 115496,  DOI:10.1016/J.CATTOD.2025.115496 .
  114. S. Awadakkam, V. Chaudhary, R. Kalagnanam, V. B. Borugadda and A. K. Dalai, Advancing hydrothermal liquefaction of Canadian forestry biomass for sustainable biocrude production: co-solvent integration, co-liquefaction, and process optimization, Sustain. Energy Fuels, 2025, 9(7), 1717–1728,  10.1039/D4SE01347F .
  115. D. Ocampo, E. A. Gómez, L. A. Ríos and G. J. Vargas, Effects of the use of acetone as co-solvent on the financial viability of bio-crude production by hydrothermal liquefaction of CO2 captured by microalgae, J. CO2 Util., 2024, 89, 102960,  DOI:10.1016/J.JCOU.2024.102960 .
  116. M. Usman, S. Cheng, S. Boonyubol and J. S. Cross, From biomass to biocrude: Innovations in hydrothermal liquefaction and upgrading, Energy Convers. Manag., 2024, 302, 118093,  DOI:10.1016/J.ENCONMAN.2024.118093 .
  117. HTL-Reactors | Task 34, Accessed: Sep. 14, 2025. [Online]. Available: https://task34.ieabioenergy.com/htl-reactors/ Search PubMed.
  118. J. L. Wagner, C. D. Le, V. P. Ting and C. J. Chuck, Design and operation of an inexpensive, laboratory-scale, continuous hydrothermal liquefaction reactor for the conversion of microalgae produced during wastewater treatment, Fuel Process. Technol., 2017, 165, 102–111,  DOI:10.1016/J.FUPROC.2017.05.006 .
  119. O. Mohan, O. A. Fakayode and A. Kumar, Biofuel production from pipeline-transported lignocellulosic biomass via hydrothermal liquefaction: Process optimization and product characterization, Biomass Bioenergy, 2025, 202, 108210,  DOI:10.1016/J.BIOMBIOE.2025.108210 .
  120. P. Tirumareddy, B. R. Patra, V. B. Borugadda and A. K. Dalai, Co-hydrothermal liquefaction of waste biomass: Comparison of various feedstocks and process optimization, Bioresour. Technol. Rep., 2024, 27, 101898,  DOI:10.1016/J.BITEB.2024.101898 .
  121. N. Jokinen, et al., Valorization potential of the aqueous products from hydrothermal liquefaction and stepwise slow pyrolysis of wood bark and hemp hurds with yields and product comparison, Bioresour. Technol. Rep., 2023, 21, 101385,  DOI:10.1016/J.BITEB.2023.101385 .
  122. J. Yip, M. Chen, Y. S. Szeto and S. Yan, Comparative study of liquefaction process and liquefied products from bamboo using different organic solvents, Bioresour. Technol., 2009, 100(24), 6674–6678,  DOI:10.1016/J.BIORTECH.2009.07.045 .
  123. Z. Liu and F. S. Zhang, Effects of various solvents on the liquefaction of biomass to produce fuels and chemical feedstocks, Energy Convers. Manag., 2008, 49(12), 3498–3504,  DOI:10.1016/J.ENCONMAN.2008.08.009 .
  124. Y. Qian, C. Zuo, J. Tan and J. He, Structural analysis of bio-oils from sub-and supercritical water liquefaction of woody biomass, Energy, 2007, 32(3), 196–202,  DOI:10.1016/J.ENERGY.2006.03.027 .
  125. H. Durak and T. Aysu, Effects of catalysts and solvents on liquefaction of Onopordum heteracanthum for production of bio-oils, Bioresour. Technol., 2014, 166, 309–317,  DOI:10.1016/J.BIORTECH.2014.05.051 .
  126. Y. Qu, X. Wei and C. Zhong, Experimental study on the direct liquefaction of Cunninghamia lanceolata in water, Energy, 2003, 28(7), 597–606,  DOI:10.1016/S0360-5442(02)00178-0 .
  127. C. Xu and T. Etcheverry, Hydro-liquefaction of woody biomass in sub- and super-critical ethanol with iron-based catalysts, Fuel, 2008, 87(3), 335–345,  DOI:10.1016/J.FUEL.2007.05.013 .
  128. M. Hupa, O. Karlström and E. Vainio, Biomass combustion technology development – It is all about chemical details, Proc. Combust. Inst., 2017, 36(1), 113–134,  DOI:10.1016/J.PROCI.2016.06.152 .
  129. B. Yan, et al., Application of optical diagnosis technology in biomass combustion, Biomass Bioenergy, 2024, 184, 107198,  DOI:10.1016/J.BIOMBIOE.2024.107198 .
  130. J. Li, M. C. Paul, P. L. Younger, I. Watson, M. Hossain and S. Welch, Characterization of biomass combustion at high temperatures based on an upgraded single particle model, Appl. Energy, 2015, 156, 749–755,  DOI:10.1016/J.APENERGY.2015.04.027 .
  131. S. A. El-Sayed, M. E. Mostafa, T. M. Khass, E. H. Noseir and M. A. Ismail, Combustion and mass loss behavior and characteristics of a single biomass pellet positioning at different orientations in a fixed bed reactor, Biomass Convers. Biorefin., 2024, 14(14), 15373–15393,  DOI:10.1007/S13399-023-03767-Z/FIGURES/7 .
  132. J. Riaza, et al., Combustion of single biomass particles in air and in oxy-fuel conditions, Biomass Bioenergy, 2014, 64, 162–174,  DOI:10.1016/J.BIOMBIOE.2014.03.018 .
  133. O. Karlström, A. Brink and M. Hupa, Time dependent production of NO from combustion of large biomass char particles, Fuel, 2013, 103, 524–532,  DOI:10.1016/J.FUEL.2012.06.030 .
  134. A. Toptas, Y. Yildirim, G. Duman and J. Yanik, Combustion behavior of different kinds of torrefied biomass and their blends with lignite, Bioresour. Technol., 2015, 177, 328–336,  DOI:10.1016/J.BIORTECH.2014.11.072 .
  135. X. Liu, M. Chen and Y. Wei, Combustion behavior of corncob/bituminous coal and hardwood/bituminous coal, Renew. Energy, 2015, 81, 355–365,  DOI:10.1016/J.RENENE.2015.03.021 .
  136. P. E. Mason, L. I. Darvell, J. M. Jones, M. Pourkashanian and A. Williams, Single particle flame-combustion studies on solid biomass fuels, Fuel, 2015, 151, 21–30,  DOI:10.1016/J.FUEL.2014.11.088 .
  137. Y. Haseli, J. A. van Oijen and L. P. H. de Goey, A detailed one-dimensional model of combustion of a woody biomass particle, Bioresour. Technol., 2011, 102(20), 9772–9782,  DOI:10.1016/J.BIORTECH.2011.07.075 .
  138. J. Zhang, et al., Synergistic effect and volatile emission characteristics during co-combustion of biomass and low-rank coal, Energy, 2024, 289, 130015,  DOI:10.1016/J.ENERGY.2023.130015 .
  139. H. Bi, C. Wang, Q. Lin, X. Jiang, C. Jiang and L. Bao, Combustion behavior, kinetics, gas emission characteristics and artificial neural network modeling of coal gangue and biomass via TG-FTIR, Energy, 2020, 213, 118790,  DOI:10.1016/J.ENERGY.2020.118790 .
  140. P. Thornley, Increasing biomass based power generation in the UK, Energy Policy, 2006, 34(15), 2087–2099,  DOI:10.1016/J.ENPOL.2005.02.006 .
  141. D. R. McIlveen-Wright, et al., A technical and economic analysis of three large scale biomass combustion plants in the UK, Appl. Energy, 2013, 112, 396–404,  DOI:10.1016/J.APENERGY.2012.12.051 .
  142. A. A. Khan, W. de Jong, P. J. Jansens and H. Spliethoff, Biomass combustion in fluidized bed boilers: Potential problems and remedies, Fuel Process. Technol., 2009, 90(1), 21–50,  DOI:10.1016/J.FUPROC.2008.07.012 .
  143. S. Rajaram, Next generation CFBC, Chem. Eng. Sci., 1999, 54(22), 5565–5571,  DOI:10.1016/S0009-2509(99)00288-2 .
  144. X. Ke, et al., Development of biomass-fired circulating fluidized bed boiler with high steam parameters based on theoretical analysis and industrial practices, J. Energy Inst., 2022, 105, 415–423,  DOI:10.1016/J.JOEI.2022.10.011 .
  145. Power from Waste: The World's Biggest Biomass Power Plants, Accessed: Jan. 14, 2025. [Online]. Available: https://www.power-technology.com/features/featurepower-from-waste-the-worlds-biggest-biomass-power-plants-4205990/?cf-view Search PubMed.
  146. OPET-organisations for the Promotion of Energy Technologies the World's Largest Biofuel CHP Plant Alholmens Kraft, Pietarsaari, Accessed: Jan. 14, 2025. [Online]. Available: https://www.tekes.fi/opet/ Search PubMed.
  147. Y. A. Begum, S. Kumari, S. K. Jain and M. C. Garg, A review on waste biomass-to-energy: integrated thermochemical and biochemical conversion for resource recovery, Environ. Sci. Adv., 2024, 3(9), 1197–1216,  10.1039/D4VA00109E .
  148. M. Volpe, E. Parvari, D. Mahajan and E. L. Hewitt, A Review of Biomass Pyrolysis for Production of Fuels: Chemistry, Processing, and Techno-Economic Analysis, Biomass, 2025, 5, 54,  DOI:10.3390/BIOMASS5030054 .
  149. E. Danso-Boateng, O.-W. Achaw, E. Danso-Boateng and O.-W. Achaw, Bioenergy and biofuel production from biomass using thermochemical conversions technologies—a review, AIMS Energy, 2022, 4, 585–585647,  DOI:10.3934/ENERGY.2022030 .
  150. H. Ge, J. Zheng and H. Xu, Advances in machine learning for high value-added applications of lignocellulosic biomass, Bioresour. Technol., 2023, 369, 128481,  DOI:10.1016/J.BIORTECH.2022.128481 .
  151. M. Meena, S. Shubham, K. Paritosh, N. Pareek and V. Vivekanand, Production of biofuels from biomass: Predicting the energy employing artificial intelligence modelling, Bioresour. Technol., 2021, 340, 125642,  DOI:10.1016/J.BIORTECH.2021.125642 .
  152. A. Kumar Sharma, P. Kumar Ghodke, N. Goyal, S. Nethaji and W. H. Chen, Machine learning technology in biohydrogen production from agriculture waste: Recent advances and future perspectives, Bioresour. Technol., 2022, 364, 128076,  DOI:10.1016/J.BIORTECH.2022.128076 .
  153. R. Wang, et al., Enhancing biomass conversion to bioenergy with machine learning: Gains and problems, Sci. Total Environ., 2024, 927, 172310,  DOI:10.1016/J.SCITOTENV.2024.172310 .
  154. J. E. Schmidt and P. Cihan, Bayesian Hyperparameter Optimization of Machine Learning Models for Predicting Biomass Gasification Gases, Appl. Sci., 2025, 15, 1018,  DOI:10.3390/APP15031018 .
  155. F. E. Yatim, et al., Blending biomass-based liquid biofuels for a circular economy: Measuring and predicting density for biodiesel and hydrocarbon mixtures at high pressures and temperatures by machine learning approach, Renew. Energy, 2024, 234, 121146,  DOI:10.1016/J.RENENE.2024.121146 .
  156. A. Butean, I. Cutean, R. Barbero, J. Enriquez and A. Matei, A Review of Artificial Intelligence Applications for Biorefineries and Bioprocessing: From Data-Driven Processes to Optimization Strategies and Real-Time Control, Processes, 2025, 13, 2544,  DOI:10.3390/PR13082544 .
  157. I. Rojek, D. Mikołajewski, A. Mroziński and M. Macko, Green Energy Management in Manufacturing Based on Demand Prediction by Artificial Intelligence—A Review, Electronics, 2024, 13, 3338,  DOI:10.3390/ELECTRONICS13163338 .
  158. O. I. Abiodun, A. Jantan, A. E. Omolara, K. V. Dada, N. A. E. Mohamed and H. Arshad, State-of-the-art in artificial neural network applications: A survey, Heliyon, 2018, 4(11), e00938,  DOI:10.1016/J.HELIYON.2018.E00938/ASSET/85B25AED-7CEE-48F9-8F25-73F74B8D991E/MAIN.ASSETS/GR6.JPG .
  159. M. khan, et al., Applications of machine learning in thermochemical conversion of biomass-A review, Fuel, 2023, 332, 126055,  DOI:10.1016/J.FUEL.2022.126055 .
  160. X. Yang, C. Yuan, S. He, D. Jiang, B. Cao and S. Wang, Machine learning prediction of specific capacitance in biomass derived carbon materials: Effects of activation and biochar characteristics, Fuel, 2023, 331, 125718,  DOI:10.1016/J.FUEL.2022.125718 .
  161. M. Madadi, E. Kargaran, S. Al Azad, M. Saleknezhad, E. Zhang and F. Sun, Machine learning-driven optimization of biphasic pretreatment conditions for enhanced lignocellulosic biomass fractionation, Energy, 2025, 326, 136241,  DOI:10.1016/J.ENERGY.2025.136241 .
  162. Y. Qiao, E. Kargaran, H. Ji, M. Madadi, S. Rafieyan and D. Liu, Data-driven insights for enhanced cellulose conversion to 5-hydroxymethylfurfural using machine learning, Bioresour. Technol., 2025, 430, 132582,  DOI:10.1016/J.BIORTECH.2025.132582 .
  163. Y. Ayub and J. Ren, Co-Pyrolysis of biomass and plastic waste: Process prediction and optimization based on Artificial Intelligence and response optimizer surrogate model, Process Saf. Environ. Prot., 2024, 186, 612–624,  DOI:10.1016/J.PSEP.2024.04.049 .
  164. N. S. Rajpurohit, P. K. Kamani, M. Lenka and C. S. Rao, Predictive modeling of product yields in microwave-assisted co-pyrolysis of biomass and plastic with enhanced interpretability using explainable AI approaches, J. Anal. Appl. Pyrolysis, 2025, 188, 107021,  DOI:10.1016/J.JAAP.2025.107021 .
  165. Z. Ullah, et al., A comparative study of machine learning methods for bio-oil yield prediction – A genetic algorithm-based features selection, Bioresour. Technol., 2021, 335, 125292,  DOI:10.1016/J.BIORTECH.2021.125292 .
  166. M. Khan, Z. Ullah, O. Mašek, S. Raza Naqvi and M. Nouman Aslam Khan, Artificial neural networks for the prediction of biochar yield: A comparative study of metaheuristic algorithms, Bioresour. Technol., 2022, 355, 127215,  DOI:10.1016/J.BIORTECH.2022.127215 .
  167. T. Onsree and N. Tippayawong, Machine learning application to predict yields of solid products from biomass torrefaction, Renew. Energy, 2021, 167, 425–432,  DOI:10.1016/J.RENENE.2020.11.099 .
  168. X. Zhou, J. Zhao, M. Chen, S. Wu, G. Zhao and S. Xu, Effects of hydration parameters on chemical properties of biocrudes based on machine learning and experiments, Bioresour. Technol., 2022, 350, 126923,  DOI:10.1016/J.BIORTECH.2022.126923 .
  169. S. Wu, J. Zhao, C. Li, X. Li, Z. Xu and H. Wang, Machine learning prediction of bio-polyol yields and hydroxyl values from acid-catalyzed liquefaction of lignocellulosic biomass, Ind. Crops Prod.Oct, 2024, 218, 119022,  DOI:10.1016/J.INDCROP.2024.119022 .
  170. S. Sezer and U. Özveren, Investigation of syngas exergy value and hydrogen concentration in syngas from biomass gasification in a bubbling fluidized bed gasifier by using machine learning, Int. J. Hydrogen Energy, 2021, 46(39), 20377–20396,  DOI:10.1016/J.IJHYDENE.2021.03.184 .
  171. A. Sakheta, T. Raj, R. Nayak, I. O'Hara and J. Ramirez, Improved prediction of biomass gasification models through machine learning, Comput. Chem. Eng., 2024, 191, 108834,  DOI:10.1016/J.COMPCHEMENG.2024.108834 .
  172. H. A. Baloch, et al., Recent advances in production and upgrading of bio-oil from biomass: A critical overview, J. Environ. Chem. Eng., 2018, 6(4), 5101–5118,  DOI:10.1016/J.JECE.2018.07.050 .
  173. F. Stankovikj, A. G. McDonald, G. L. Helms, M. V. Olarte and M. Garcia-Perez, Characterization of the Water-Soluble Fraction of Woody Biomass Pyrolysis Oils, Energy Fuels, 2017, 31(2), 1650–1664,  DOI:10.1021/ACS.ENERGYFUELS.6B02950/SUPPL_FILE/EF6B02950_SI_001.PDF .
  174. M. Garcia-Perez, S. Wang, J. Shen, M. Rhodes, W. J. Lee and C. Z. Li, Effects of Temperature on the Formation of Lignin-Derived Oligomers during the Fast Pyrolysis of Mallee Woody Biomass, Energy Fuels, 2008, 22(3), 2022–2032,  DOI:10.1021/EF7007634 .
  175. Y. Han, et al., Hydrotreatment of pyrolysis bio-oil: A review, Fuel Process. Technol., 2019, 195, 106140,  DOI:10.1016/J.FUPROC.2019.106140 .
  176. J. G. M. Ucella-Filho, et al., Antiviral potential of bio-oil from Citrus sinensis waste wood as a therapeutic approach against COVID-19, J. Clean. Prod., 2024, 447, 141583,  DOI:10.1016/J.JCLEPRO.2024.141583 .
  177. A. D. Abd-Elnabi, E. Abdel Fattah El-sawy and E. M. El-Adawy, Insecticidal effects of the fast pyrolysis bio-oil against Spodoptera littoralis and Aphis gossypii insect pests, J. Asia Pac. Entomol., 2024, 27(2), 102237,  DOI:10.1016/J.ASPEN.2024.102237 .
  178. M. Ahmad, et al., Biochar as a sorbent for contaminant management in soil and water: A review, Chemosphere, 2014, 99, 19–33,  DOI:10.1016/J.CHEMOSPHERE.2013.10.071 .
  179. J. S. Cha, et al., Production and utilization of biochar: A review, J. Ind. Eng. Chem., 2016, 40, 1–15,  DOI:10.1016/J.JIEC.2016.06.002 .
  180. Y. Chen, X. Zhang, W. Chen, H. Yang and H. Chen, The structure evolution of biochar from biomass pyrolysis and its correlation with gas pollutant adsorption performance, Bioresour. Technol., 2017, 246, 101–109,  DOI:10.1016/J.BIORTECH.2017.08.138 .
  181. H. Kazemi Shariat Panahi, et al., A comprehensive review of engineered biochar: Production, characteristics, and environmental applications, J. Clean. Prod., 2020, 270, 122462,  DOI:10.1016/J.JCLEPRO.2020.122462 .
  182. Y. Xie, et al., A critical review on production, modification and utilization of biochar, J. Anal. Appl. Pyrolysis, 2022, 161, 105405,  DOI:10.1016/J.JAAP.2021.105405 .
  183. D. V. Cuong, et al., A critical review on biochar-based engineered hierarchical porous carbon for capacitive charge storage, Renew. Sustain. Energy Rev., 2021, 145, 111029,  DOI:10.1016/J.RSER.2021.111029 .
  184. S. Nambyaruveettil, L. Ali, M. B. Beg, A. Khaleel and M. Altarawneh, Novel biochar-based hybrid support catalyst for the selective hydrogenation of 1,3-butadiene: A complete conversion with exceptional deactivation resistance, Chem. Eng. J. Adv., 2025, 23, 100779,  DOI:10.1016/J.CEJA.2025.100779 .
  185. S. Packialakshmi, B. Anuradha, K. Nagamani, J. Sarala Devi and S. Sujatha, Treatment of industrial wastewater using coconut shell based activated carbon, Mater. Today Proc., 2023, 81(2), 1167–1171,  DOI:10.1016/J.MATPR.2021.04.548 .
  186. V. Ahuja, et al., Biochar: Empowering the future of energy production and storage, J. Anal. Appl. Pyrolysis, 2024, 177, 106370,  DOI:10.1016/J.JAAP.2024.106370 .
  187. Y. Zhang, J. Zhang, K. Song and W. Lv, Potential of biochar derived from three biomass wastes as an electrode catalyzing oxygen reduction reaction, Environ. Pollut. Bioavailab., 2022, 34(1), 42–50,  DOI:10.1080/26395940.2021.2022538 .
  188. S. Nambyaruveettil, L. Ali, M. B. Beg, A. Khaleel and M. Altarawneh, A novel Ni–zeolite–biochar green catalyst: optimization of composition and preparation strategy revealed by TPR, TPD, and XPS, Mater. Sci. Eng. B, 2025, 321, 118566,  DOI:10.1016/J.MSEB.2025.118566 .
  189. D. W. Cho, S. H. Cho, H. Song and E. E. Kwon, Carbon dioxide assisted sustainability enhancement of pyrolysis of waste biomass: A case study with spent coffee ground, Bioresour. Technol., 2015, 189, 1–6,  DOI:10.1016/J.BIORTECH.2015.04.002 .
  190. T. Sizmur, T. Fresno, G. Akgül, H. Frost and E. Moreno-Jiménez, Biochar modification to enhance sorption of inorganics from water, Bioresour. Technol., 2017, 246, 34–47,  DOI:10.1016/J.BIORTECH.2017.07.082 .
  191. B. Igliński, W. Kujawski and U. Kiełkowska, Pyrolysis of Waste Biomass: Technical and Process Achievements, and Future Development—A Review, Energies, 2023, 16, 1829,  DOI:10.3390/EN16041829 .
  192. C. S. Damian, Y. Devarajan, T. Raja and R. Jayabal, A comprehensive review of biomass pyrolysis for hydrogen production in India, Process Saf. Environ. Prot., 2024, 190, 646–662,  DOI:10.1016/J.PSEP.2024.07.034 .
  193. T. Kan, V. Strezov and T. J. Evans, Lignocellulosic biomass pyrolysis: A review of product properties and effects of pyrolysis parameters, Renew. Sustain. Energy Rev., 2016, 57, 1126–1140,  DOI:10.1016/J.RSER.2015.12.185 .
  194. Y. Zhang, et al., A review of biomass pyrolysis gas: Forming mechanisms, influencing parameters, and product application upgrades, Fuel, 2023, 347, 128461,  DOI:10.1016/J.FUEL.2023.128461 .
  195. M. Yu, et al., A 4E analysis of a novel coupling process of syngas purification and CO2 capture, transcritical CO2 power and absorption refrigeration, Chem.–Eng. J., 2022, 445, 136757,  DOI:10.1016/J.CEJ.2022.136757 .
  196. P. Straka and J. Cihlář, Conversion of Biomass to Energy-Rich Gas by Catalyzed Pyrolysis in a Sealed Pressure Reactor, Processes, 2025, 13, 1692,  DOI:10.3390/PR13061692 .
  197. M. B. Brands, P. Beuel, F. Torres-Rivera, R. Beckmüller, M. S. Ayoub and P. Stenzel, Optimization of biomethane production from purified biowaste pyrolysis gas: A comparative techno-economic assessment, Renew. Energy, 2025, 245, 122865,  DOI:10.1016/J.RENENE.2025.122865 .
  198. C. Guo, K. T. V. Rao, Z. Yuan, S. (Quan) He, S. Rohani and C. Charles) Xu, Hydrodeoxygenation of fast pyrolysis oil with novel activated carbon-supported NiP and CoP catalysts, Chem. Eng. Sci., 2018, 178, 248–259,  DOI:10.1016/J.CES.2017.12.048 .
  199. L. Ali, M. S. Kuttiyathil and M. Altarawneh, Catalytic upgrading of the polymeric constituents in Covid-19 masks, J. Environ. Chem. Eng., 2022, 10(1), 106978,  DOI:10.1016/j.jece.2021.106978 .
  200. N. Chaihad, S. Karnjanakom, A. Abudula and G. Guan, Zeolite-based cracking catalysts for bio-oil upgrading: A critical review, Resour. Chem. Mater., 2022, 1(2), 167–183,  DOI:10.1016/J.RECM.2022.03.002 .
  201. M. M. Rahman, R. Liu and J. Cai, Catalytic fast pyrolysis of biomass over zeolites for high quality bio-oil – A review, Fuel Process. Technol., 2018, 180, 32–46,  DOI:10.1016/J.FUPROC.2018.08.002 .
  202. P. Soldatos, et al., Conversion of beechwood organosolv lignin via fast pyrolysis and in situ catalytic upgrading towards aromatic and phenolic-rich bio-oil, Sustain. Chem. Environ., 2024, 6, 100107,  DOI:10.1016/J.SCENV.2024.100107 .
  203. V. Chaerusani, et al., In-situ catalytic upgrading of bio-oils from rapid pyrolysis of torrefied giant miscanthus (Miscanthus x giganteus) over copper-magnesium bimetal modified HZSM-5, Appl. Energy, 2024, 353, 122110,  DOI:10.1016/J.APENERGY.2023.122110 .
  204. A. M. Venezia, X-ray photoelectron spectroscopy (XPS) for catalysts characterization, Catal. Today, 2003, 77(4), 359–370,  DOI:10.1016/S0920-5861(02)00380-2 .
  205. A. Yaghi, L. Ali, T. Shittu, M. S. Kuttiyathil, A. Khaleel and M. Altarawneh, Hydrodeoxygenation of Vapor Anisole over Nickel/Cobalt and Alumina/Zeolite Supported Catalysts, Catal. Surv. Asia, 2024, 28(1), 48–57,  DOI:10.1007/S10563-023-09409-8/METRICS .
  206. O. Ismail, et al., Hydro-deoxygenation of waste biomass pyrolysates on cobalt-sulfided catalyst for the production of BTX fuels, Case Stud. Chem. Environ. Eng., 2024, 9, 100734,  DOI:10.1016/J.CSCEE.2024.100734 .
  207. L. Ali, et al., Catalytic upgrading of bio-oil from halophyte seeds into transportation fuels, J. Bioresour. Bioprod., 2023, 8(4), 444–460,  DOI:10.1016/J.JOBAB.2023.10.002 .
  208. S. De, B. Saha and R. Luque, Hydrodeoxygenation processes: Advances on catalytic transformations of biomass-derived platform chemicals into hydrocarbon fuels, Bioresour. Technol., 2015, 178, 108–118,  DOI:10.1016/J.BIORTECH.2014.09.065 .
  209. A. Dimitriadis, G. Meletidis, U. Pfisterer, M. Auersvald, D. Kubička and S. Bezergianni, Integration of stabilized bio-oil in light cycle oil hydrotreatment unit targeting hybrid fuels, Fuel Process. Technol., 2022, 230, 107220,  DOI:10.1016/J.FUPROC.2022.107220 .
  210. S. Li, Y. Wu, M. U. Dao, E. N. Dragoi and C. Xia, Spotlighting of the role of catalysis for biomass conversion to green fuels towards a sustainable environment: Latest innovation avenues, insights, challenges, and future perspectives, Chemosphere, 2023, 318, 137954,  DOI:10.1016/J.CHEMOSPHERE.2023.137954 .
  211. M. S. Kuttiyathil, et al., Catalytic upgrading of pyrolytic bio-oil from Salicornia bigelovii seeds for use as jet fuels: Exploring the ex-situ deoxygenation capabilities of Ni/Ze catalyst, Bioresour. Technol. Rep., 2023, 22, 101437,  DOI:10.1016/J.BITEB.2023.101437 .
  212. K. Du, et al., Interface enhanced catalytic hydrodeoxygenation of vanillin as a bio-oil model using Ni-based catalysts, Chem.–Eng. J., 2025, 520, 166037,  DOI:10.1016/J.CEJ.2025.166037 .
  213. J. X. Xie, et al., Highly efficient conversion of lignin bio-oil and derived phenols to cyclohexanols over low-loading Ni/C catalyst, Fuel, 2024, 371, 132030,  DOI:10.1016/J.FUEL.2024.132030 .
  214. A. R. K. Gollakota, et al., Catalytic hydrodeoxygenation of bio-oil and model compounds - Choice of catalysts, and mechanisms, Renew. Sustain. Energy Rev., 2023, 187, 113700,  DOI:10.1016/J.RSER.2023.113700 .
  215. N. T. Machado, et al., Upgrading/Deacidification of Bio-Oils by Liquid–Liquid Extraction Using Aqueous Methanol as a Solvent, Energies, 2024, 17(11), 2713,  DOI:10.3390/EN17112713/S1 .
  216. B. Zhao, et al., Influence of extraction solvents on the recovery yields and properties of bio-oils from woody biomass liquefaction in sub-critical water, ethanol or water–ethanol mixed solvent, Fuel, 2022, 307, 121930,  DOI:10.1016/J.FUEL.2021.121930 .
  217. D. Chiaramonti, et al., Development of emulsions from biomass pyrolysis liquid and diesel and their use in engines—Part 1 : emulsion production, Biomass Bioenergy, 2003, 25(1), 85–99,  DOI:10.1016/S0961-9534(02)00183-6 .
  218. D. Chiaramonti, et al., Development of emulsions from biomass pyrolysis liquid and diesel and their use in engines—Part 2: tests in diesel engines, Biomass Bioenergy, 2003, 25(1), 101–111,  DOI:10.1016/S0961-9534(02)00184-8 .
  219. S. Bhat, V. B. Borugadda and A. K. Dalai, Emulsification of bio-crude produced from agricultural waste via hydrothermal liquefaction process, Fuel, 2021, 305, 121602,  DOI:10.1016/J.FUEL.2021.121602 .
  220. J. W. Chong, N. G. Chemmangattuvalappil, K. Muthoosamy and S. Thangalazhy-Gopakumar, Stable diesel and bio-oil emulsion through solvent extraction and ultrasonic aided emulsification, Process Saf. Environ. Prot., 2024, 190, 616–624,  DOI:10.1016/J.PSEP.2024.07.038 .
  221. X. L. Li, et al., Green degradation of Caragana korshinskii to valuable bio-oils in supercritical CO2-ethanol: Evaluation of CH and O moieties, J. Anal. Appl. Pyrolysis, 2023, 174, 106154,  DOI:10.1016/J.JAAP.2023.106154 .
  222. B. Rezvani, Novel techniques of pretreatments and post-treatments for bio-oil upgrading: A comprehensive review, Biomass Bioenergy, 2025, 201, 108086,  DOI:10.1016/J.BIOMBIOE.2025.108086 .
  223. Y. Fan, Y. Han, J. Zhu, Y. Chen, Y. Cai and W. Zhao, Production optimization of refined bio-oil through plasma coupling catalysis and composite zeolite effect of Al-SBA-15/HZSM-5, Fuel, 2023, 347, 128494,  DOI:10.1016/J.FUEL.2023.128494 .
  224. Z. P. Fu, et al., Selective hydrodeoxygenation of guaiacol and bio-oil to cycloalkanes over Ni-based catalysts supported on altered Hβ zeolite, Mol. Catal., 2025, 576, 114932,  DOI:10.1016/J.MCAT.2025.114932 .
  225. H. Hu, et al., Unveiling the enhancement of catalytic activity and stability by single-atom Mo anchored to cationic vacancy for the catalytic hydrodeoxygenation of lignin to naphthenes, Chem.–Eng. J., 2025, 505, 159303,  DOI:10.1016/J.CEJ.2025.159303 .
  226. S. Cheng, L. Wei, J. Julson and M. Rabnawaz, Upgrading pyrolysis bio-oil through hydrodeoxygenation (HDO) using non-sulfided Fe-Co/SiO2 catalyst, Energy Convers. Manag., 2017, 150, 331–342,  DOI:10.1016/J.ENCONMAN.2017.08.024 .
  227. A. R. Ardiyanti, A. Gutierrez, M. L. Honkela, A. O. I. Krause and H. J. Heeres, Hydrotreatment of wood-based pyrolysis oil using zirconia-supported mono- and bimetallic (Pt, Pd, Rh) catalysts, Appl. Catal., A, 2011, 407(1–2), 56–66,  DOI:10.1016/J.APCATA.2011.08.024 .
  228. Y. Luo, V. K. Guda, E. B. Hassan, P. H. Steele, B. Mitchell and F. Yu, Hydrodeoxygenation of oxidized distilled bio-oil for the production of gasoline fuel type, Energy Convers. Manag., 2016, 112, 319–327,  DOI:10.1016/J.ENCONMAN.2015.12.047 .
  229. A. H. Hamid, et al., Transformation of levoglucosan into liquid fuel via catalytic upgrading over Ni-CeO2 catalysts, Mol. Catal., 2023, 547, 113382,  DOI:10.1016/J.MCAT.2023.113382 .
  230. Y. Xu, et al., In situ hydrogenation of model compounds and raw bio-oil over Raney Ni catalyst, Energy Convers. Manag., 2015, 89, 188–196,  DOI:10.1016/J.ENCONMAN.2014.09.017 .
  231. A. Sanna, T. P. Vispute and G. W. Huber, Hydrodeoxygenation of the aqueous fraction of bio-oil with Ru/C and Pt/C catalysts, Appl. Catal., B, 2015, 165, 446–456,  DOI:10.1016/J.APCATB.2014.10.013 .
  232. O. Ismail, et al., Selective formation of fuel BXT compounds from catalytic hydrodeoxygenation of waste biomass over Ni-decorated beta-zeolite, Bioresour. Technol. Rep., 2023, 24, 101616,  DOI:10.1016/J.BITEB.2023.101616 .
  233. S. Xiu and A. Shahbazi, Bio-oil production and upgrading research: A review, Renewable Sustainable Energy Rev., 2012, 16(7), 4406–4414,  DOI:10.1016/j.rser.2012.04.028 .
  234. A. Carrasco Díaz, L. Abdelouahed, N. Brodu, V. Montes-Jiménez and B. Taouk, Upgrading of Pyrolysis Bio-Oil by Catalytic Hydrodeoxygenation, a Review Focused on Catalysts, Model Molecules, Deactivation, and Reaction Routes, Molecules, 2024, 29, 4325,  DOI:10.3390/MOLECULES29184325 .
  235. Z. Tian, et al., Hydrodeoxygenation of guaiacol as a model compound of pyrolysis lignin-oil over NiCo bimetallic catalyst: Reactivity and kinetic study, Fuel, 2022, 308, 122034,  DOI:10.1016/J.FUEL.2021.122034 .
  236. Y. Wu, et al., Catalytic Hydrodeoxygenation of Pyrolysis Volatiles from Pine Nut Shell over Ni-V Bimetallic Catalysts Supported on Zeolites, Catalysts, 2025, 15(5), 498,  DOI:10.3390/CATAL15050498/S1 .
  237. X. Cao, J. Zhao, F. Long, X. Zhang, J. Xu and J. Jiang, Al-modified Pd@mSiO2 core-shell catalysts for the selective hydrodeoxygenation of fatty acid esters: Influence of catalyst structure and Al atoms incorporation, Appl. Catal., B, 2022, 305, 121068,  DOI:10.1016/J.APCATB.2022.121068 .
  238. Y. Deng, et al., Hydrodeoxygenation of phenolic compounds and lignin pyrolysis oil for the production of cyclohexanes over core-shell Ru@SiO2-ZrO2 catalyst with high stability, Fuel, 2026, 404, 136219,  DOI:10.1016/J.FUEL.2025.136219 .
  239. S. Shao, et al., A review on the application of non-thermal plasma (NTP) in the conversion of biomass: Catalyst preparation, thermal utilization and catalyst regeneration, Fuel, 2022, 330, 125420,  DOI:10.1016/J.FUEL.2022.125420 .
  240. I. M. Smarte Anekwe and Y. Makarfi Isa, Unlocking catalytic longevity: a critical review of catalyst deactivation pathways and regeneration technologies, Energy Adv., 2025, 4(9), 1075–1113,  10.1039/D5YA00015G .
  241. L. Pinard and C. Batiot-Dupeyrat, Non-thermal plasma for catalyst regeneration: A review, Catal. Today, 2024, 426, 114372,  DOI:10.1016/J.CATTOD.2023.114372 .
  242. X. Li, H. Zhang, S. Shao, Z. Lv, S. Ge and Y. Cai, Direct non-thermal plasma regeneration of deactivated HZSM-5 for catalytic pyrolysis of rape straw, J. Anal. Appl. Pyrolysis, 2021, 157, 105209,  DOI:10.1016/J.JAAP.2021.105209 .
  243. Y. Shen, Biomass pretreatment for steam gasification toward H2-rich syngas production – An overview, Int. J. Hydrogen Energy, 2024, 66, 90–102,  DOI:10.1016/J.IJHYDENE.2024.04.096 .
  244. H. S. Choi, Y. S. Choi and H. C. Park, Fast pyrolysis characteristics of lignocellulosic biomass with varying reaction conditions, Renew. Energy, 2012, 42, 131–135,  DOI:10.1016/J.RENENE.2011.08.049 .
  245. J. Y. Zhu, Physical pretreatment - Woody biomass size reduction - For forest biorefinery, ACS Symp. Ser., 2011, 1067, 89–107,  DOI:10.1021/BK-2011-1067.CH004 .
  246. Z. E. Zadeh, A. Abdulkhani, O. Aboelazayem and B. Saha, Recent Insights into Lignocellulosic Biomass Pyrolysis: A Critical Review on Pretreatment, Characterization, and Products Upgrading, Processes, 2020, 8, 799,  DOI:10.3390/PR8070799 .
  247. S. Zhang, et al., Liquefaction of Biomass and Upgrading of Bio-Oil: A Review, Molecules, 2019, 24, 2250,  DOI:10.3390/MOLECULES24122250 .
  248. E. A. Silveira, et al., Biofuel from agro-industrial residues as sustainable strategy for CO2 mitigation: Statistical optimization of pequi seeds torrefaction, Energy Convers. Manag., 2024, 304, 118222,  DOI:10.1016/J.ENCONMAN.2024.118222 .

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.