Next Article in Journal
Optimal Sizing of Photovoltaic/Energy Storage Hybrid Power Systems: Considering Output Characteristics and Uncertainty Factors
Next Article in Special Issue
Perspectives on the Development of Technologies for Hydrogen as a Carrier of Sustainable Energy
Previous Article in Journal
A Study of the Decoupling of Economic Growth from CO2 and HFCs Emissions in the EU27 Countries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Influence of Catalyst on the Yield and Quality of Bio-Oil for the Catalytic Pyrolysis of Biomass: A Comprehensive Review

by
Md Sumon Reza
1,2,*,
Zhanar Baktybaevna Iskakova
1,
Shammya Afroze
3,
Kairat Kuterbekov
3,
Asset Kabyshev
3,*,
Kenzhebatyr Zh. Bekmyrza
3,
Marzhan M. Kubenova
4,
Muhammad Saifullah Abu Bakar
2,
Abul K. Azad
2,
Hridoy Roy
5 and
Md Shahinoor Islam
5,6,*
1
Research Institute of New Chemical Technologies, Faculty of Natural Sciences, L.N. Gumilyov Eurasian National University, Astana 010008, Kazakhstan
2
Faculty of Integrated Technologies, Universiti Brunei Darussalam, Jalan Tungku Link, Gadong BE1410, Brunei
3
Faculty of Physics and Technical Sciences, L.N. Gumilyov Eurasian National University, Astana 010008, Kazakhstan
4
Faculty of Transport and Energy, L.N. Gumilyov Eurasian National University, Astana 010008, Kazakhstan
5
Department of Chemical Engineering, Bangladesh University of Engineering and Technology, Dhaka 1000, Bangladesh
6
Department of Textile Engineering, Daffodil International University, Dhaka 1341, Bangladesh
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(14), 5547; https://doi.org/10.3390/en16145547
Submission received: 24 May 2023 / Revised: 17 July 2023 / Accepted: 20 July 2023 / Published: 22 July 2023

Abstract

:
In the modern world, as the population rises and fossil fuel supplies decline, energy demands continue to rise. Moreover, the use of fossil fuels harms the ecology, contributing to pollution and global warming. In order to overcome these difficulties, several approaches are revealed, such as the utilization of biomass as a renewable source of energy. Studies revealed that biomass can be converted into bioenergy via several thermal conversion processes, like pyrolysis, gasification, and torrefaction. Pyrolysis is the most convenient process to obtain three different types of biofuels (biochar as a solid, bio-oil as a liquid, and syngas as a gas). The biofuels produced in this process are normally lower in quality and cannot be used directly as fuel because they contain many undesirable components. Catalytic pyrolysis is one of the best processes to upgrade the quality of biofuels. Several varieties of catalysts are used in the catalytic pyrolysis process (ex situ and in situ). Due to stable operating conditions, both catalytic and non-catalytic pyrolysis procedures produce biochar that has a consistent output. Meanwhile, the effects of catalysts in the catalytic pyrolysis process considerably enhance the quality and quantity of bio-oils and syngas. By removing the unwanted oxygenated and nitrogenous components, the bio-oils produced through the catalytic pyrolysis method have a higher calorific value, reduced viscosity, and improved stability. Many researchers have looked at ways to increase the rate of pyrolysis, whereas a few have focused on maximizing the effects of the factors in order to improve the efficiency of catalytic pyrolysis. This review addresses the impact of catalysts on the catalytic pyrolysis of biomass to enhance the quality of the bio-oils in great detail. Machine learning and techno-economic analysis were investigated, as well as the future potential of the catalytic pyrolysis method for the generation of bio-oil.

1. Introduction

One of the most important resources for people is energy, which forges a close connection between nature and life. Energy needs are growing daily since there are no alternatives for them in the social and economic growth of a nation [1,2,3]. The primary energy source is the combustion of fossil fuels, which increases emissions of greenhouse gases, mostly carbon dioxide (CO2) [4,5]. The implications of worldwide climate change and environmental obligations make renewable energy sources necessary today. Also, the need for energy from sources of renewable energy that should be both ecologically and economically advantageous is being driven by the depletion of fossil fuel supplies [6,7]. These problems can be solved using renewable energy sources, including biomass, wind, solar, geothermal, and tidal, where there is great potential for converting biomass to biofuels [8,9].
Agricultural waste, municipal trash, food waste, grasses, wood waste, and other types of waste are among the primary sources of biomass [10]. Dry plant material, also known as lignocellulosic biomass, is a plentiful and inexpensive source of renewable energy. Biomass can be converted into bioenergy through thermochemical or biochemical conversion processes. In these conversion phases, the thermochemical conversion implicates heat, and the biochemical conversion implicates enzymes [11]. In these methods, thermochemical conversion is the most promising and efficient way to convert biomass into bioenergy when compared to the biochemical process [12].
Among the major thermochemical conversion methods (pyrolysis, gasification, and combustion) [13], pyrolysis is more appropriate because of its benefits for storage, transportation, and solicitation flexibility [14]. Biochar, bio-oil, and syngas are the three primary byproducts of the pyrolysis of biomass [15]. It was also found from the literature that excess vapor cracking occurs at high operating temperatures to boost gas production. Low temperatures increase char yield, but temperatures in the 400–600 °C range produce the highest bio-oil output in various reactors [16]. Biochar is often used to produce activated carbon (AC), improve soil, generate heat, and capture CO2 [17]. AC is ideal for catalytic applications, including water and air filtration, due to its enormous surface area, considerable porosity, and structural durability [18]. Bio-oils can be purified and utilized as low-grade fuel for heating and diesel engines [19]. Syngas is upgraded and refined, before being utilized to generate energy and power in fuel cells and gas generators [20].
The majority of unwanted components (volatile acids, oxygenated compounds, and water) that are present in bio-oils after pyrolysis lower the heating value of oil. The catalytic pyrolysis procedures used to produce upgraded bio-oils are of great interest to researchers [21]. Catalysts are materials that speed up chemical reactions while remaining unchanged. In catalytic pyrolysis, catalysts can decrease oxygenated and nitrogenous components, enhance calorific value, reduce viscosity, and improve the solidity of the bio-oils [22]. Catalysts for the catalytic pyrolysis process most frequently include zeolites, metal oxides, clay, dolomite, and activated carbon [23].
The kind of catalyst employed, the structure of the biomass, and operational parameters like temperature, catalyst-to-biomass ratio, reactor type, particle size, heating rate, and residence time, affect the yield and quality of the bio-oil generated during catalytic pyrolysis [24,25]. Normally, the yield of bio-oil produced by catalytic pyrolysis is somewhat lower than that of non-catalytic pyrolysis [26]. According to earlier research, different catalysts have distinct potentialities in the catalytic pyrolysis process [27]. The release of oxygen and carbon oxides (COx) during catalytic pyrolysis eventually affects the yield and quality of bio-oils for the deoxygenation reactions (dehydration, decarbonylation, and decarboxylation) [28].
Pyrolysis has great promise, but commercialization is still difficult to achieve. As an outcome, it is becoming increasingly important to fully comprehend its dynamics to enable process build up. However, biomass pyrolysis is a complicated, sluggish, and expensive process. In spite of these difficulties, machine learning (ML) has come to be seen as a potential technique for assisting and advancing pyrolysis research. The handling of vast amounts of data, flexibility in complicated systems, and computing efficiency are some of the benefits of using ML. Due to its enormous potential, ML may be successfully used to simulate pyrolysis processes and produce predictions that are equivalent to those made by conventional models [29]. In addition, a vibrant field of research focuses on the effectiveness of machine learning in process optimization and control, product yield prediction, real-time monitoring, and techno-economic analysis (TEA) during catalytic pyrolysis. The incorporation of ML appears to be the most effective strategy for expediting research and development and expanding the scope of the catalytic pyrolysis of biomass.
The primary focus of this study is on state-of-the-art catalytic pyrolysis using lignocellulosic biomass for the generation of bio-oil for environmental sustainability. This review summarizes the impacts of various biomass types, catalyst types, and other parameters that significantly influence the yield and quality of bio-oils. The present study also includes machine learning and techno-economic analyses of the catalytic pyrolysis process. It provides a fundamental understanding of present developments and the direction of future research.

2. Biomass and Thermochemical Conversion of Biomass

2.1. Biomass

Biomass is the waste product of plants or animals that cannot be used as food or feed. Plants, food waste, crops and forestry waste, agro-industrial waste, garbage, municipal solid waste, and biodegradable industrial solid waste are examples of different biomass sources (Figure 1) [30]. Woody biomass is made up of trees, bushes, and shrubs, whereas herbaceous biomass refers to plants with non-woody stems, such as those that yield grains and cereals. As a possible source of renewable energy, biomass—also known as a feedstock or solid biofuel—can reduce greenhouse gas emissions more than fossil fuels. If these biomass materials are not adequately handled, transported, or disposed of, they will pose an environmental threat by producing greenhouse gases [31,32].
Hemicellulose, which makes up 20–40% of the biomass, is followed by cellulose, which makes up 40–60% of the biomass, and lignin, which makes up 10–25% of the biomass [33,34]. Comparing cellulose to hemicellulose and lignin, during the catalytic pyrolysis process, cellulose produced the greatest amount of aromatic hydrocarbons [35]. Figure 2 illustrates the structure and major constituents of lignocellulosic biomass.

2.1.1. Cellulose

Cellulose is the crystalline and amorphous form of a glucose molecule with the formula (C6H10O5)n, which is bonded through hydrogen bonds and Van der Waals forces [37]. It is extremely insoluble in water and has a high resistance to depolymerization. In biomass, this is the most important part of the cell wall. Crystalline cellulose has a high resistance to degradation compared to amorphous cellulose [38]. More volatile matter may be produced by biomass with increased cellulose content to produce significant amounts of bio-oil for the generation of biofuel [39]. The bulk of olefins generated during the pyrolysis process come from biomass with a greater cellulose or hemicellulose content than from feedstock with a higher lignin content [40].

2.1.2. Hemicellulose

Hemicellulose is the amorphous and heterogeneous branch of polymers with the overall formula (C5H8O4)n. It is a component of the cell wall with a lack of strength and a lower molecular weight than cellulose [37]. It is a pronged glucose or xylose polymer with lactose, xylose, arabinose, fucose, glucose, mannose, and glucuronic acid as supplements [38]. The degree of polymerization makes it amorphous, readily hydrolyzable, and unable to combine with cellulose. Hemicellulose can support the pyrolysis process to produce volatile and gaseous materials due to its lower molecular weight and shorter polymer chains [41]. In the catalytic pyrolysis process, hemicellulose is one of the most effective elements to produce higher amounts of aromatic hydrocarbons [42].

2.1.3. Lignin

Lignin is a 3D (three-dimensional), cross-connected polymer made up of phenylpropane units joined together by hydroxyl and methoxy groups. It protects hemicellulose and cellulose from enzymes, spores, and bacteria by being contained in the outer layer of the tissue [43]. Lignin is made up of three simple structural units: guaiacyl type (G-type), hydroxyl phenyl (H-type), and syringyl type (S-type) [44]. Boiling water, salt, or any other alkali solvent does not cause lignin to dissolve [45]. Lignin has a higher carbon content than cellulose and hemicellulose, as well as a lower oxygen concentration [46]. It decomposes at a slower rate in pyrolysis since it is often used to generate biochar [47]. Lignin is the least thermally degradable component during pyrolysis, which produces the phenolic compounds in bio-oils [48].

2.1.4. Organic Extractives

Biomass extractives include alkaloids, mineral oils, glycosides, fats, gums, pectins, resins, starches, phenolics, proteins, waxes, and terpenes [49]. They are made up of both water-soluble (sugars and proteins) and ethanol-soluble (waxes and chlorophyll) constituents [50]. Extractives can be inorganic or organic. The biomass with a higher amount of extractives is in pyrolysis as inorganic extractives produce low-quality bio-oils, and organic extractives cause clogging in the reactor [51].

2.1.5. Inorganic Matter/Ash

Biomass also holds many minerals or inorganic elements that produce ash after complete combustion [52]. The elements of ash in the biomass are mostly potassium (K), silicon (Si), sodium (Na), calcium (Ca), phosphorus (P), iron (Fe), manganese (Mn), aluminum (Al), chromium (Cr), zinc (Zn), and magnesium (Mg) [49]. According to Antal et al. [53], the majority of the ash in char is made up of alkali metals, which have a catalytic influence on the yield, reactivity, and characteristics of pyrolysis products. Inorganics may accumulate on the surface of the catalyst or inside its pores throughout the catalytic pyrolysis process, which might lead to the catalyst losing its effectiveness. Since the inorganics may be isolated before the catalytic reactions for ex situ procedures, the inorganics have a greater impact on in situ catalytic pyrolysis processes than on ex situ processes [54].

2.2. Thermochemical Conversion of Biomass

Thermochemical conversion is the method of transforming biomass into solid, liquid, and gaseous products. There are several types of thermochemical processes, like pyrolysis, gasification, and combustion [55]. Figure 3 demonstrates the thermochemical conversion processes and their major products [56].
Which form of thermal conversion will take place is virtually entirely controlled by heat and oxygen. The basic premise behind these ideas is that when oxygen is added, the biomass will transition from pyrolysis to gasification, followed by combustion with the further addition of oxygen. Pyrolysis is the most basic form of conversion. It only needs a small amount of heat and no oxygen to begin breaking down biomass molecules. Gasification is the intermediate stage of conversion. Compared to pyrolysis, it takes higher heat, but more crucially, it also needs a tiny amount of oxygen. Combustion is the ultimate stage of conversion. It combusts to provide the system with all the oxygen it can take and enough heat to set it off. Therefore, pyrolysis produces just a small quantity of heat, followed by gasification, which produces more heat, and combustion, which produces the maximum heat [57].

2.2.1. Combustion

Fuel and oxygen are burnt during combustion, which is an exothermic reaction that generates heat, carbon dioxide, and water [58]. The steam turbine uses combustion to turn the blades positioned on the shaft linked to the generator by producing hot steam from the burning of fuel. This process converts heat energy to kinetic energy, mechanical energy, and ultimately electrical energy [59]. However, because it causes global warming by releasing carbon dioxide into the atmosphere, the use of fossil fuels is not entirely environmentally friendly. When burning biomass, additional polluting gases can be produced, including potassium chloride (KCl (g), K2Cl2 (g)), hydrogen chloride, HCl (g), and sodium chloride, NaCl (g) [60,61].

2.2.2. Gasification

Gasification is a thermochemical conversion process that converts carbonaceous materials, such as biomass or coal, into fuel gas when the gasifying agent, air, is present. The gas produced, which is composed of H2, CH4, CO2, CO, and N2, is known as “syngas.” Additionally, syngas produces a little amount of char, ashes, and tar [62]. With this approach, waste may be managed efficiently, limiting the impact of landfills and greenhouse gas emissions [63]. Syngas can be an effective fuel for solid oxide fuel cells (SOFCs), enhancing and raising the value of biomass as a renewable energy source. Gas cleaning is important in the gasification process for use in SOFCs. Various forms of catalysts are also used in gasifiers to upgrade the products [64,65].

2.2.3. Pyrolysis

Pyrolysis is the process of thermally breaking down biomass into smaller molecules at higher temperatures and in an inert atmosphere [56]. The primary outcomes of pyrolysis are biochar, bio-oil, and syngas. Slow pyrolysis is concerned with biochar processing, while fast pyrolysis is mostly used to generate bio-oils and syngas [66]. During the decomposition process, the volatiles are released as vapor, and the non-volatiles are converted into solid biochar. The condensable vapors are referred to as bio-oil, which is also identified as pyrolysis oil, bio-crude oil, and biofuel oil, whereas the non-condensable gases are known as syngas, biogas, and bio-syngas [67]. When biomass is heated, volatile substances are liberated from the particles, forming non-condensable gases [68]. Syngas is produced predominantly by gasification. It is mostly employed in catalysis. Carbon monoxide and hydrogen are the two most crucial syngas ingredients for chemical synthesis. It may be used to produce methanol and ammonia, which lessen the need for natural gas in each case [69]. Meanwhile biogas is the anaerobic digestion or fermentation of biodegradable materials, such as biomass, manure, sewage, municipal trash, green waste, plant material, and energy crops [70]. Typically, carbon dioxide and methane make up their compositions.

3. Pyrolysis of Biomass

3.1. Types of Pyrolysis

There are four main categories of pyrolysis: slow, intermediate, fast, and flash pyrolysis (Figure 4). The circumstances and parameters for each form of pyrolysis vary, depending on factors including heating temperature, residence duration, heating rate, and sample size [71]. The operating conditions and typical product yields for each kind of pyrolysis are shown in Table S1, where the production of bio-oil increases from the slow to the fast pyrolysis process [72,73].

3.1.1. Slow Pyrolysis

In the slow pyrolysis process, the temperature used is between 300 and 500 °C, with a very long solid residence time of several minutes to several hours at a lower heating rate. The main products of this process are mainly biochar and low-value energy products [74].

3.1.2. Intermediate Pyrolysis

In intermediate pyrolysis, the reaction temperature ranges from 400 to 600 °C, and the residence period can reach several minutes. The sample can be either coarse, shredded, or finely ground [14]. Intermediate pyrolysis is effective for bio-oil production from grasses, straws, and agricultural residues [75]. In this process, high production of bio-oil can be achieved at a moderate temperature (500 °C) with an increasing reaction time [76].

3.1.3. Fast Pyrolysis

Fast pyrolysis has a reaction temperature between 400 °C and 1000 °C, a high heating rate of up to 1000 °C/s, and a brief residence time of less than 2 s [72]. Small particle size, fast heating rate, and moderate pyrolysis temperatures (400 and 650 °C) are the common characteristics of the fast pyrolysis process [73]. In this technique, the production of bio-oil might increase to 80%, while the biochar and syngas yields are very low [77]. The kinetic behavior of fast pyrolysis is perplexing because of the complexity of chemical reactions. The mineral composition of the feedstock for the catalytic activity also has an impact on yields [78]. Catalytic fast pyrolysis is a potential technique that might enable the conversion of biomass into portable fuels. The catalytic fast pyrolysis of biomass has not been used commercially despite its benefits in terms of simplicity and economic efficiency [79].

3.1.4. Flash Pyrolysis

With a heating rate of more than 1000 °C/s and a dwell time of under 0.5 s, flash pyrolysis takes place at temperatures that are comparatively high (900–1300 °C) [80]. The major product from flash pyrolysis is bio-oil (>75 wt%) [81]. It is a complex process, as with an upper heating rate and a smaller residence period, the biomass must be pre-treated and made into a fine powder (105–250 μm) before pyrolysis to ensure efficient heat transfer [82].

3.2. Reactors for the Pyrolysis Processes

Several types of pyrolysis technologies are available depending on the required end products (syngas, bio-oil, or biochar) [81]. The design of the pyrolysis reactor is the most important factor influencing product distribution and quality. Each technology contains specific working principles and an experimental setup [83]. According to the mobility of the biomass inside the reactor, the technologies are classified as fixed-bed reactors, auger/screw reactors, and fluidized bed reactors. Figure 5 depicts the available pyrolysis technologies. The benefits and drawbacks of adopting various pyrolysis technologies are listed in Table S2, with a fixed-bed reactor being the most efficient and practical reactor on a lab size [84,85,86,87].

3.2.1. Fixed-Bed Reactor

The reactor is positioned in a vertical or horizontal position in the fixed-bed reactor. Vertically, the solid particles move down via a shaft and contact the gas of the product stream, whereas horizontally, the particle is fixed inside the reactor [89]. Biochar production via slow pyrolysis using a fixed-bed reactor is simple to perform. During carbonization, the biomass is put into a reactor for many hours or even days to produce charcoal as the major product. This is the simplest form of carbonization, where only a chimney is required to combust the biomass. It is low in costs and simple to perform [90].

3.2.2. Auger/Screw Reactor

The auger reactor, sometimes referred to as a screw reactor, is made up of a revolving helical screw that serves as a conveyor to transport the feedstock through the heated reactor [91]. A rotating screw ensures uniform mixing of the feedstock, which increases heat-transfer efficiency [86]. An auger reactor can be made up of single or twin screws. A twin-screw auger reactor provides agitation of the biomass and simple conveyance, allowing better mixing and, hence, better heat transfer [92]. It is a continuous pyrolysis technology and can be used for biomass of different sizes and shapes [86].

3.2.3. Fluidized Bed reactor

One of the most widely used and commercially accessible methods for creating bio-oil is the fluidized bed reactor, which is well suited for fast pyrolysis [93]. The existence of turbulent gas flow and quick circulation in the reactor allows it to have excellent heat-transfer efficiency and superior temperature control, making it a better alternative than a fixed-bed reactor [94]. The batch-dense fluidized bed reactor, which is typically comprised of a refractory steel cylinder tube and a set of electric furnaces, is one of the most often used fluidized bed reactors [95]. The process is highly reactive and has a lower residence time [84].

3.2.4. Microwave Reactor

One of the most current research topics in the pyrolysis process is a microwave reactor, which eventually transfers energy to molecules or atoms through microwave heat. Normally, microwave cavity ovens are used to pyrolyze the biomass in this process. Inert gas is used as the carrier gas, which needs to flow continuously through the reactor to maintain an oxygen-free environment [88]. For the slow pyrolysis process, microwave reactors are a successful way to produce valuable compounds from biomass because they have several benefits, including effective heat transmission, enhanced chemical reactivity, and control of the heating process exponentially [96]. As a result, microwave reactors can process a variety of biomasses and industrial wastes with high yields of syngas and bio-oil [97].

3.2.5. Other Reactors

In addition, a number of other reactors were used to study the catalytic pyrolysis of biomass, including dual fluidized bed reactors [98], conical-spouted bed reactors [99], circulating fluidized bed reactors [100], drop-tube quartz reactors [101], and V-shaped downer reactors [102].

3.2.6. Impacts of Pyrolysis and Reactors on the Enhancement of Bio-Oil Quality

Ignited earlier, numerous studies have already demonstrated that the most affordable, easiest, and most scalable process for converting biomass and other feedstock into energy and other useful byproducts is pyrolysis [103]. The physicochemical characteristics, quality, and yield of pyrolyzed products (e.g., bio-oil) can all be greatly impacted by pyrolysis process parameters [104,105]. The performance and quality of the product are significantly influenced by several factors, including raw materials, reactor type, temperature, additives, catalysts, residence time, and pressure [106]. Bio-oil has a particularly difficult set of physicochemical characteristics due to the abundance of oxygenated organic functionalities that are present in it, and this bio-oil is transformed through extreme secondary upgrading into, for example, transportation fuels [107].
Fast pyrolysis is thought to be the most effective method of all these thermal conversion techniques for producing liquid fuel, such as bio-oil, at its highest potential. This pyrolysis is capable of producing up to 75% of bio-oil, which may be utilized immediately in a variety of applications or upgraded to become an energy carrier [108]. With a brief dwelling period of around 2–10 s and higher heating rates, fast pyrolysis is an effective method for turning biomass into bio-oil in an inert environment at medium temperatures between 400 and 600 °C. It emerged that this system could generate a maximum output of bio-oil of 78.07 wt% at the bed temperature of 480 °C by employing a circulating fluidized bed reactor via low-cost bio-oil fast pyrolysis [109]. Another study revealed that the fast pyrolysis of S. japonica in a fixed-bed reactor yielded a maximum bio-oil output of 40.91 wt% at 350 °C and a sweeping-gas velocity of 300 mL/min [110]. From the foregoing considerations, it can be inferred that although the fixed-bed reactor is more frequently employed, fast pyrolysis with a fluidized bed reactor produces the highest quantity of bio-oil output.
However, compared to bio-oil produced by fast pyrolysis, intermediate pyrolysis-produced bio-oil has better physical and chemical characteristics, including a low oxygen concentration, greater oil miscibility with fossil fuels, and a high calorific value [111]. Fast pyrolysis yields more bio-oil (60–75%) than intermediate pyrolysis (35–50%) in terms of yield [71,75,112,113,114]. Compared to fast pyrolysis oil (35–40%), intermediate pyrolysis bio-oil contains less oxygen (8–33%). Bio-oil with a high oxygen concentration is unstable and contains lower energy content [115]. In comparison to fast pyrolysis oil, intermediate pyrolysis oil has lower viscosity and miscibility due to the low oxygen level, making it easier to blend with biodiesel and hence easier to use in conventional engines. Consequently, intermediate pyrolysis is a viable choice for processing waste with little to no pretreatment while still producing high-quality products. It also has the potential to co-produce adsorption bio-oil, demonstrating that waste conversion is a technology that is both resource- and environment-friendly.

3.3. Products from the Pyrolysis Process

The main outcomes of the pyrolysis process are shown in Figure 6, where the solid is biochar, the liquid is bio-oil, and the gas is syngas.

3.3.1. Biochar

Biochar, also known as charcoal or biocharcoal, is a carbonaceous substance with a higher carbon content [116]. The breakdown of lignin, which breaks weak bonds and forms more dense, stable structures from primary and secondary pyrolysis events, results in the formation of biochar [117]. It can be used to produce heat energy by containing more carbon (63–91%) and less oxygen [118]. Biochar production is aided by the higher content of ash due to inorganic substances [119]. Devolatilization assists in the formation of biochar with improved porosity and reactivity as biomass is pyrolyzed at high temperatures [66]. Biochar is used to diversify the arena due to its available functional groups, inert nature, and capability to repossess liquid or gaseous particles [120]. Biochar has recently demonstrated remarkable potential in the manufacture of activated carbon, which is one of the most adaptable adsorbents for a wide range of specialized applications [121]. Biochar is also useful as a solid fuel and fertilizer [114].

3.3.2. Bio-Oil

Pyrolysis is one of the most efficient processes for producing bio-oils from biomass, with up to 70% efficiency [122]. It is a solvent derived from the condensable vapors of a complex mixture of acids, phenols, ketones, alcohols, aromatic hydrocarbons, carbohydrates, esters, ethers, amines, oxygenated compounds, and water [123]. Typically, bio-oils from the pyrolysis process mostly appear dark brown, are viscous/non-viscous, and have a strong, distinctive smell [124]. The production of bio-oils is mostly determined by the heat breakdown of cellulose [125]. After refining, bio-oils can be utilized directly as fuel for boilers, furnaces, or engines [126]. They are regarded as an acceptable fuel for energy savings, environmental support, and rural socioeconomic concerns [127]. Bio-oils can also be used for wood flavorings, preservatives, lubricants, resins, and adhesive synthesis, in addition to biofuels [128].
The two stages of bio-oil production are the organic phase and the aqueous phase, with the organic phase usually having greater energy levels than the aqueous phase [129].
The viscosity, acidity, and thermal stability of pyrolytic bio-oil are greater, resulting in bio-oil deterioration over time [130]. For appropriate applications, it is crucial to improve the quality of bio-oil. Currently, three forms of bio-oil updating processes are available:
  • Physical methods with solvents;
  • Catalytic methods with catalysts;
  • Non-catalytic methods without catalysts.
Among the above processes, the catalytic methods show higher performance due to the better quality of the bio-oils [114].

3.3.3. Syngas

Depending on the feedstock, the gases produced during the pyrolysis process are referred to as syngas, biogas, or bio-syngas. H2, CH4, CO2, CO, and hydrocarbons are the principal constituents of the non-condensable gases. The production of gas grows with rising temperature as a result of the secondary breaking of the vapors [66]. Syngas can be utilized as a direct fuel in fuel cells, especially in microtubular solid oxide fuel cells, as a combustion fuel in manufacturing, as a heat source in the pyrolysis process, or as fuel in gas turbines to produce electricity [121,131]. The gas products from the pyrolysis process offer benefits over charcoal and bio-oil since they do not need any pricey upgrading operations [114].

3.4. Factors Influencing the Pyrolysis Process and Products

Both the pyrolysis procedure and the pyrolysis yields were significantly influenced by a number of different variables: temperature, heating rate, residence time, biomass particle size, presence of the catalyst, etc. [132].

3.4.1. Temperature

The temperature has a significant impact on the yield and effectiveness of the pyrolysis processes and the products [133]. Normally, if the pyrolysis temperature increases, the production of biochar decreases by enhancing the pyrolytic vapors [134]. In general, the yield of bio-oils was achieved for the pyrolysis temperature of 400 to 550 °C; after that (>550 °C), the production is reduced by converting the tars and vapors into non-condensable gases via secondary cracking reactions, such as decarboxylation, deoxygenation, and dehydrogenation [135,136].

3.4.2. Heating Rate

The product yield in the pyrolysis process is significantly influenced by the heating rate. The pyrolytic vapors from biomass rise due to a greater rate of breakdown and produce more gases than charcoal as a result of the increased heating rate [137]. For the pyrolysis of 14 plants (Wiregrass, Inkberry, Yaupon, Fetterbush, Swamp bay, Wax Myrtle, Longleaf Pine Foliage, Water Oak, Little Bluestem, Live Oak, Dwarf Palmetto, Saw Palmetto, Sparkleberry, Darrow’s Blueberry), an escalation of the heating rate from 0.5 °C/s to 180 °C/s increases the tar and light gas yield and decreases the biochar yield [138]. Increasing heating rates leads to higher secondary reactions, which produce more gaseous products [139]. When the heating rate increased from 500 to 700 °C/min for the pyrolysis of sawdust, it was discovered that the yield of bio-oil increased by 8%, but when the rate increased from 700 °C/min to 1000 °C/min, there was no discernible difference in the production of bio-oil [140]. Similar outcomes were obtained while pyrolyzing cottonseed cake, where the bio-oil yield ranged from 26 wt% for a heating rate of 5 °C/min and 35 wt% for a heating rate of 300 °C/min. There was little change in the yield of bio-oils for the increase in heating rate from 300 to 700 °C/min [141].

3.4.3. Vapor Residence Time

The length of stay, which significantly affects output yield, is the length of time that a substance persists in a medium or location. The amount of time it takes for the vaporized gas inside the reactor to escape is known as the vapor residence time [142]. Since preventive secondary reactions take place in the reactors by eliminating organic vapors during shorter residence durations, more bio-oil is produced [143]. The intermediate pyrolysis process produced bio-oil with greater quality for a short residence time than the fast pyrolysis technique in terms of physical and chemical attributes [144]. On the other hand, with a long vapor residence time, the slow pyrolysis process provides a lower yield of bio-oils [145]. It was found that the yield of syngas increased with a higher residence time [146]. In the pyrolysis (at 525 °C) of raw sorghum bagasse, the bio-oil yields were reduced from 75 to 57% with an increase in vapor residence time from 0.2 s to 0.9 s, increasing the yield of biochar and syngas [136].

3.4.4. Biomass Particle Size

Particle size influences product yield; as particle size increases, gas yield decreases, while char and tar yields rise due to the lower surface area and higher thermal resistance [147]. In the pyrolysis of smaller particles, the temperature is constant throughout each particle, which produces more liquids and gases by full conversion. On the other hand, for bigger particles, the inside is heated slowly, resulting in a stay in the solid phase due to lower temperatures producing higher amounts of solids. For a higher amount of bio-oils, smaller particles are highly preferable [148]. As a consequence, the production of biochar and syngas decreases, but the output of bio-oil increases due to the moderate particle size. The Chinese tallow tree produced the highest amount of bio-oil (46.20%) when the biomass particles were between 1.4 and 2.4 mm in size and the pyrolysis temperature was 550 °C [149].

3.4.5. Presence of Catalysts

The liquids, gases, and solids generated during pyrolysis are greatly influenced by the presence of a catalyst. Few catalysts were achieved to increase the solid products, whereas others enhanced the liquids and gaseous products [71]. In comparison to the non-catalytic process, the production of bio-oil reduced from 36.70 to 34.72% in the ex situ catalytic pyrolysis of invasive Pennisetum purpureum grass with activated carbon, whereas the output of syngas increased from 35.90 to 37.52%. This was because of the catalytic effect of the catalytic and pyrolytic vapors [26].
The pyrolysis procedure can be accomplished either thermally (traditionally) or catalytically. It is an endothermic procedure [150]. In contrast to the catalytic approach, which involves using an external agent to help the pyrolysis process, the thermal route involves the process occurring without the aid of any external agents [151]. Conventional pyrolysis produces liquid byproducts that have long carbon chains with low octane values, wax, other residues, and impurities, necessitating additional refining to transform them into useful fuel products [152]. To address such challenges, catalytic pyrolysis of biomass is a powerful thermochemical conversion technique that may be utilized to enhance the quality of the liquid or bio-oil by boosting the number of lighter fractions in the oil and lowering the impurity concentration.

4. Catalytic Pyrolysis of Biomass

4.1. Catalytic Pyrolysis

Catalysts are substances that accelerate chemical processes without undergoing permanent chemical changes. They can open up new pathways by decreasing the energy required for a chemical reaction to happen under particular circumstances, such as pressure and temperature [153]. At the operating temperature range of 400 to 650 °C, catalysts are added to catalytic pyrolysis to increase the yield of the products [154]. When related to non-catalytic pyrolysis, the heavy components are broken down into smaller and lighter components, such as hydrocarbons, overall reducing oxygenated compounds during the cracking reaction [155]. There are two classifications of catalysts used in pyrolysis, namely primary and secondary catalysts [156]. The types of catalysts and the pyrolysis setup for the catalytic pyrolysis of biomass are described in Table 1.
The prime catalysts are those that are combined with the biomass in the primary chamber before the pyrolysis process (in situ), while the secondary catalysts are those that are located in the secondary chamber after the pyrolysis process (ex situ). In the secondary process, the combustion of syngas or bio-oil takes place for energy or heat generation [173]. Sometimes, catalytic pyrolysis can also take place in the reactor chamber without mixing biomass and the catalyst [174].

4.1.1. Arrangement of Catalysts

In situ or ex situ catalytic pyrolysis can be categorized according to where the catalysts are located inside the reactor [175]. The thermochemical conversion of biomass is carried out by in situ (primary) and ex situ (secondary) catalytic processes, as shown in Figure 7 [176,177]. By choosing the right catalysts and operating conditions, both procedures are necessary to transform oxygenated and heavy substances into chemically stable molecular products.

In Situ

In the in situ catalytic pyrolysis procedure, catalysts and biomass particles are mixed during the pyrolysis procedure to enhance the reactions. Catalysts not only participate in reactions but also serve as a heating medium. After being released from the biomass particles, the pyrolysis vapors directly react with the nearby catalysts in the in situ phase. This proximity between the biomass and catalyst speeds up the conversion of larger molecules. To provide effective deoxygenation of the vapors, there is a need for a large catalyst-to-biomass ratio [22]. The disadvantage of this process is that the biomass and catalyst have to operate under the same pyrolysis conditions. It is also tough to separate the biochar from the catalyst after pyrolysis [178]. The yield and quality of bio-oil were negatively impacted by the catalyst surface’s abnormally high coke generation, according to research [96].

Ex Situ

Catalysts are positioned in the same or other reactors in the ex situ catalytic pyrolysis process, downstream of the biomass. With various configurations of catalysts, it can be set up in a single bed, two phases, or more in sequence [179]. This method provides more stability and controllability for having various reaction conditions (combinations of catalysts, different temperatures, and variations in the gaseous condition), which can be dissimilar from the main biomass reactor. In addition, the ex situ process makes it possible to recover and recycle catalysts without the difficult isolation of catalysts that occurs in the in situ process [180]. As a consequence, both the end product and the intermediate product may be accurately tracked and tailored in line with research and industry demands after each procedure [22]. Ex situ pyrolysis is constrained by financial constraints due to the need for an extra catalytic reactor. Lower oxygen concentrations in bio-oils were discovered in an earlier investigation utilizing ex situ pyrolysis, even though deoxygenating activities were not seen during the in situ process [181].

4.1.2. Affecting Factors of Catalyst

Catalyst Position

The position of the catalyst (in situ and ex situ setups) plays an important role in the catalytic pyrolysis of biomass. When utilizing bubbling fluidized bed (BFB), auger, or pyro probe reactors, better yields of bio-oil, aromatics, phenols, gas, and CO are obtained in in situ mode with ZSM-5 (Zeolite Socony Mobil–5) compared to ex situ results. For the auger reactor, an in situ arrangement is preferable to an ex situ one for removing oxygen. When the BFB reactor was utilized, however, the opposite tendency was noted. Ex situ mode seems to yield better results for the deoxygenation of bio-oils in a fixed-bed catalytic pyrolysis reactor. Ex situ arrangements are preferred because they stop coking or metal accumulation from deactivating the catalyst by preventing it from happening on the surface of the catalyst. As a result, it is tough to conclude that the mode is superior because the findings are inconclusive [54].

Ratio of Catalyst and Biomass

The “biomass-to-catalyst ratio”, also known as “weight turnover”, is the proportion of biomass or feed weight to catalyst weight. In the realm of catalytic pyrolysis, it is well known that catalysts can lose their activation properties when exposed to prolonged reactions or large amounts of feed materials because coke can build up on the catalyst’s surface and in its micropores, blocking the active sites [54]. The product yield is increased by a larger catalyst-to-feed ratio by preventing undesirable heat breakdown processes in the homogeneous phase. According to the research, increasing the catalyst ratio increased liquid yield while decreasing solid yield [182]. More than half of the Brønsted acidity and microporosity was lost during a weight turnover of 2.0 for catalytic pyrolysis of cellulose in a bubbling fluidized bed with ZSM-5. As a result, the aromatase-producing activity was significantly reduced [183]. Because the weight of the catalyst utilized is frequently significantly larger than that of the biomass feedstock, the inverse of this parameter, namely the catalyst-to-feed ratio, is occasionally used. Research suggests that increasing the catalyst-to-feed ratio might increase the generation of aromatic hydrocarbons [184].

Catalyst Deactivation

The decrease in catalytic activity over time is known as catalyst deactivation. All catalysts are deactivated for poisoning, coking, and sintering [185]. Chemisorption of substances that block the active sites of catalysts cause poisoning to occur. Chemical and physical therapy are both effective ways to remove poisons. The active sites of the catalyst are blocked or altered by vapors that are deposited on its surface, leading to coking or fouling. By keeping the higher heating rate of the process constant and raising the catalyst-to-biomass ratio, it is easier to reduce it [186]. To decrease the surface area of the catalyst sites, tiny crystals start to form inside the catalyst, which then start the process of sintering. The amount of time the catalyst is exposed to is determined by the high-temperature stream [187]. The deactivation of the catalysts is the primary downside of the catalytic pyrolysis method [188].

Catalyst Bed Temperature

In the process of catalytic pyrolysis of biomass, the temperature of the catalyst influences the rates of thermal and catalytic reactions like cracking, dehydration, decarbonylation, decarboxylation, and aromatization. These reactions affect the yields of bio-oil and gas, the productivity and selectivity of aromatic hydrocarbons and coke, as well as the deoxygenation pathways. It is possible to synthesize aromatic hydrocarbons with good selectivity and yields by raising the catalyst temperature [189]. The catalytic pyrolysis procedure is also impacted by the temperature of the catalysts. It was found that increasing the catalyst bed temperature from 450 to 600 °C caused the bio-oil production to drop as the amount of syngas increased [190]. The catalyst temperature also influences coke formation or catalyst deactivation. As catalyst temperature rises, the coke yield falls [191]. The output of coke decreased from 12.0 wt% at 400 °C to 7.8 wt% at a 600 °C catalyst temperature when rice husk was catalytically pyrolyzed in a fluidized bed reactor with a downstream fixed bed of an HZSM-5 catalyst [192].

5. Effect of Catalyst on the Catalytic Pyrolysis Process

5.1. Catalysts

There are several advantages to using a catalyst, but there are two main ones. First, the product benefits because the catalyst may increase the amount and quality of the desired end product by enhancing the formation of aromatics in the oil, which is accomplished by altering the process parameters. Additionally, the process is improved by adding a catalyst, which increases the overall process efficiency of the pyrolysis and energy performance. Next, the process is improved by lowering process temperatures and residency times, enhancing cracking and decomposition reactions, promoting isomerization reactions, and enhancing product selectivity [193]. Several catalysts are used in the catalytic pyrolysis process: zeolite, natural catalysts (dolomite and clay), activated carbon, metal oxides, etc. [194].

5.1.1. Zeolite

Zeolite is a defined microporous form of crystalline aluminosilicates. It has a tetrahedral shape and is acidic, with channels and cavities connecting it to form a porous structure. Tetrahedral zeolites have a Si or Al central configuration with oxygen at the corners [195]. It has various acidities and an amorphous structure generated from the Brønsted and Lewis acid sites [196]. However, zeolite-based catalysts are ineffective against macromolecules in the procedure and have a slight denitrogenation effect [197]. Whilst synthetic zeolites like H-MFI, H-ZSM-5, H-BEA, and H-FAU (or H-X and H-Y single bond), and their modified forms are often used for industrial chemical processes, zeolites like chabazite and mordenite may be found naturally in rocks [198]. The petroleum industry has used ZSM-5 widely as a catalyst because of its shape selectivity, good pore size with steric hindrance, heat stability, and solid acidity [199].
Zeolite can be used as a hydrocracking catalyst due to the following: (i) greater acidity resulting in greater cracking activity; (ii) better thermal/hydrothermal stability; (iii) better naphtha selectivity; (iv) better resistance to nitrogen and sulfur compounds; (v) low coke-forming tendency; and (vi) easy renewability [200]. Nevertheless, when artificial zeolite is utilized as the catalyst, bio-oil production is reduced. Although the bio-oil output is reduced, it enhances the aromatic component of the bio-oil; as a result, once it is refined, it may be utilized as energy [201].
In Table 2, along with the results, the yield of the products from the catalytic pyrolysis of rice husk is shown. Compared to non-catalytic pyrolysis, the production of bio-oil drops from 39.61% to 38.29%, while the output of charcoal and syngas rises from 18.47% to 21.62% using a Brunei rice husk ash (BRHA) catalyst. In Table 2, aromatic hydrocarbons increase in zeolite but decrease in other catalysts. Phenol components also increase in zeolite and BRHA but decrease in other catalysts. It was found that each catalyst has its chemical selectivity, and zeolite has an influence on producing more aromatic hydrocarbons and phenols, which can be converted into biofuels after they are refined [158].

5.1.2. Natural Catalyst

Clay and dolomite are examples of natural catalysts. Clay is mostly composed of alumina, silica, and magnesia [202,203]. The chemical compound dolomite, with the formula CaMg(CO3)2, is composed of calcium, magnesium, and carbonate [204]. Dolomite is used in the catalytic pyrolysis of biomass to increase the quality of the bio-oil by oxygenating oxygenates or catalytically splitting heavy organic materials into lighter fractions [205]. It was demonstrated that dolomite is effective in removing oxygen from bio-oil, resulting in a decreased O concentration and O/C ratio in the improved bio-oil [206]. It was shown that mineral clays can accelerate the catalytic cracking of heavy compounds present in the pyrolysis liquid [202].
By breaking down and reforming the high-molecular-weight organic components during the catalytic pyrolysis of biomass, dolomite demonstrated remarkable effectiveness as a catalyst [207]. When using dolomite as a catalyst in the pyrolysis process of the tulip tree, the bio-oil components C5-C11 increase, and the ratio of H2/CO produced in syngas also increases, where it can be further treated and used as a liquid fuel [203]. Table 3 illustrates the effects of using natural catalysts (sand and dolomite) as catalysts in the pyrolysis of tulip trees. It was found that the catalytic pyrolysis of biomass using sepiolite (clay) could improve the yield of bio-oil by up to 46% regarding the non-catalytic method [208]. Another type of clay, named bentonite (aluminum phyllosilicate), showed an increase in bio-oil production of 48% in the pyrolysis of peat [209].
The literature reported that the utilization of dolomite as a catalyst would result in less bio-oil (34.26%) compared to the utilization of sand as a catalyst (49.03%) being produced during the pyrolysis process of tulip tree biomass. With a rise in syngas output, bio-oil likewise falls. Dolomite, which further breaks down the heavy components, might nevertheless enhance the quality of the oil. When the temperature rises, the breakdown of the heavier components accelerates [203].

5.1.3. Activated Carbon

Activated carbon is a substance that contains rich carbon and has a high degree of porosity and a high surface area. Due to its strong structure and surface area, it can be used to adsorb gas and further crack the components into smaller components. Both physical and chemical-activation processes can be used to transform the biochar product, which is produced during the pyrolysis of biomass, into activated carbon [26]. Activated carbon is a substance that contains a high amount of carbon and can be used as a catalyst to crack the bigger components into smaller components [210]. Utilizing activated carbon as a catalyst decreased bio-oil output while enhancing syngas yield, while biochar yield remained nearly unchanged. This is due to the reforming mechanism that occurs when a sample degrades with the AC catalyst, which causes additional cracking reactions [211]. Bio-oil production decreases as the temperatures rise due to components cracking into low-molecular-weight compounds [212].
From Table 4, it can be seen that using a catalyst decreases the bio-oil production while increasing the syngas yield and the biochar yield remains almost the same. Further reducing bio-oil output and increasing syngas yield are achieved by increasing the catalyst-to-cellulose ratio. However, it is also illustrated that the class of bio-oil was upgraded, while phenol production was increased when using a catalyst and increasing the catalyst-to-cellulose ratio. The author suggested that the reduction in bio-oil production was due to the reforming process, in which the sample underwent degradation and the volatile materials that flowed through the activated carbon led to further cracking reactions [211].
In Table 4, it is shown that an experiment was conducted on Douglas fir sawdust pellets in the pyrolysis procedure to engage in a different type of activated carbon, whose source was bituminous coal, wood, and lignite coal. In the absence of a catalyst, bio-oil yield was 45.20 wt%, syngas was 11.80 wt%, biochar was 43.00 wt%, and phenol components were 2.540 wt%. When activated carbon was introduced, the bio-oil yield and biochar production declined, while syngas production increased. However, the phenolic components were increasing significantly. This was because other components were further broken down into smaller components, where they were mostly converted into phenolic components [213].
According to the latest research, the yield of the final products during the pyrolysis process can be increased by employing activated carbon as a catalyst. Furans, ketones, ethers, and phenols, for example, can be increased in bio-oil by up to 87% when activated carbon is used as a catalyst. Activated carbon can also increase the development of hydrocarbon or aromatic composites, where a rise in pH value and high heating value can upgrade the class of bio-oil [214]. Furthermore, the breakdown of components into low-molecular-mass compounds also results in a decrease in bio-oil production as a consequence of temperature increases [212].
The co-pyrolysis of the lignin and soapstock was studied, and non-catalytic co-pyrolysis showed higher yields of bio-oil and biochar with lower gas production. The major components showed low aromatics, phenolics, and other oxygenates with high aliphatic content. While activated carbon as a catalyst was introduced, it decreased the bio-oil yields and increased the gas yield, and the major components showed an improvement in aromatics and aliphatic compounds, with very few phenolic and other oxygenated compounds. Increases in the feedstock-to-catalyst ratio will, however, result in higher yields of bio-oil and lower yields of gas, which will result in a reduction in aromatic components and an increase in aliphatic components [212].

5.1.4. Metal Oxide

Metal oxides are crystalline solids composed of a metal cation and an oxide anion [215]. Several types of metal oxides were used in catalytic pyrolyzes, like magnesium oxides (MgO), aluminum oxides (Al2O3), zinc oxides (ZnO), calcium oxides (CaO), titanium oxides (TiO2), and others [152,216]. Most of the time, metal oxides are frequently utilized in a variety of processes as heterogeneous catalysts. Metal oxides display redox qualities as a result of their polyvalent characteristics or specific acid-base characteristics, which may be useful for the catalytic pyrolysis of biomass to yield more stable products [217]. By using catalytic deoxygenation, oxygen from pyrolytic vapor is removed as CO2, CO, and H2O. Because of the carbon transfer for the gas product, the formation of water, and the carbon on the surface of the catalyst, the total yield of the liquid and organic phases is lower [163]. Metal oxides can be classified as acidic (Al2O3, SiO2, SiO2-Al2O3), alkaline (salts of potassium, sodium, magnesium, and calcium), or transitional (MnO2, CeO2, Fe2O3, ZnO, and MgO) metal oxides depending on their properties [152].
Metal oxides typically react with water to form bases or with acids to form salts [218]. By keeping volatiles in tar, these catalysts can prevent their further breakdown into gaseous products [219]. Oxides of acidic metals also have an impact on oxygenate synthesis. Al2O3, for instance, facilitates the breakdown of macromolecular methoxyphenol and sugar during the catalytic pyrolysis of sawdust [220]. Alkaline metal oxides are inexpensive and available in huge numbers, and they are extensively employed in the catalytic pyrolysis of biomass. Alkaline metal oxides may concurrently remove acidic functional groups from bio-oil by ketylation and aldol condensation to reduce corrosivity and effectively lower the oxygen content in bio-oil [152]. In the catalytic pyrolysis of biomass, the output of both bio-oil and compound composition is enhanced by the use of transition metal oxides. It was found that the levels of carboxylic acid and oxygen decreased in the bio-oil produced by catalyzing the pyrolysis of various oxides [221].
Table 5 explains the effects of non-catalytic and catalytic pyrolysis and was converted to graphics for easy visualization. Once the catalyst was added, the production of bio-oil and bio-gas increased, but the yield of solids declined. The job of the catalyst during the pyrolysis of biomass is to lower the number of oxygenated molecules in the bio-oil. The catalyst then further cracks the oxygenated substance through a dehydration and decarboxylation process, which produces water and carbon dioxide [222].
In Table 5, it is seen that the use of the catalyst in palm empty fruit bunches affects the product yield, where bio-oil and syngas yields decrease for CaO and MgO while ZnO increases. However, as the loading catalyst increases, the bio-oil yield in CaO decreases, while that in ZnO and MgO increases. The phenolic compounds decrease as the catalyst is used and with increases in catalyst loading. However, in ZnO, an increase in the loading effect of the catalyst does not change the phenol component. The anhydrosugar components decrease by breaking down into smaller components that increase the furan components. The acetic acid in CaO as a catalyst decreases, while it increases in ZnO and MgO since CaO is a basic oxide [223].
In the pyrolysis of rice husk, ZnO is utilized as a catalyst; the effect of temperature is that the production of biochar decreases while the yield of bio-oil and syngas increases. It also describes how the loading catalyst affects yields, showing how a rise in the loading catalyst enhances syngas production while reducing bio-oil output and slightly lowering biochar yield. In the pyrolysis of rice husk, the quality of bio-oil shows that non-catalytic bio-oil was 14.479%, while in catalytic bio-oil with zinc oxide, the bio-oil quality was 27.412%, consisting of light components with <120 g/mol, including alkanes, alkenes, styrene, and alkyl phenol [169].
Hence, from these studies, it can be deduced that different catalysts have different effects on the yield of the product, depending on the chemical selectivity of the catalysts.

5.1.5. Mixture of Catalysts

Some researchers conduct pyrolysis experiments with a mixture of catalysts to find the compatibility needed to generate a high yield of bio-oil. The purpose of this research is to find synergistic effects. Some of the mixture catalysts that were investigated are Ni/C, Ce/AC, MgO/ZSM-5, and SiC/ZSM-5 [224,225].
The co-pyrolysis of soybean straw and soapstock was investigated using SiC-ZSM-5 as the catalyst, and it was found that bio-oil decreases when using SiC and ZSM-5 individually as catalysts compared to no catalysts, but when the catalysts of SiC and ZSM-5 are mixed, the bio-oil increases slightly compared to when using the catalysts individually. When utilizing a catalyst or a catalyst combination in comparison to not using a catalyst, the phenoxy compound in the chemical composites of the bio-oil drops. When ZSM-5 is used as a catalyst, especially when compared with no catalyst, aromatic compounds considerably rise and somewhat decrease when ZSM-5 and SiC are combined [226].
Catalytic pyrolysis of Jatropha curcas residue was conducted by Kaewpengkrow et al. [227] and found that in the non-catalytic reaction, most of the constituents are the carboxylic acid group, while the acid group significantly decreases by introducing the catalysts. By using activated carbon as the catalyst, around 59.55% of aliphatic hydrocarbons are produced. When activated carbon was mixed with other metals, the aliphatic hydrocarbon components were decreased by increasing the aromatic hydrocarbons and N-compounds. For the biomass-to-catalyst ratio of 1:5, the aromatic hydrocarbon was significantly increased, and the oxygenating and acid components were decreased. For the feedstock-to-catalyst ratio of 1:10, the catalyst further decreased the mixture of oxygenating and acid components and increased the aliphatic and aromatic hydrocarbons.
Pyrolysis of acid-washed Cladophora glomerata was conducted, and Table 6 describes the comparison of non-catalytic and catalytic pyrolysis by using Zn/activated carbon. Biochar and syngas yields increased in catalytic pyrolysis from 34% to 37% and 20% to 28%, respectively, while bio-oil yields decreased from 46% to 35%. Oxygenates and acid compounds in bio-oil decreased when Zn/activated carbon was introduced, while nitrogenates, aromatics, and aromatic nitrogenates increased. It can be assumed that using Zn/activated carbon as a catalyst has selectivity for aromatics and nitrogenates [228]. A combination of K3PO4 and clinoptilolite or bentonite enhanced bio-oil output during the catalytic pyrolysis of switch grass in comparison to using a single catalyst [229].

5.2. Catalytic Effect on Bio-Oil

Before being employed in biofuels or other applications, the generated bio-oil has to be defined. It is characterized by following standards such as those of the American Society of Testing and Materials methods. The pH, water content, density, elemental analysis (CHNSO), ash content, higher heating value, viscosity, and other properties of bio-oil are being examined [230].

5.2.1. pH of the Bio-Oil

As bio-oil cannot be utilized as a transportation fuel if it is acidic, the pH of the bio-oil must be determined. As the pyrolytic temperature increases during the non-catalytic pyrolysis of pine wood chips, the pH level of the bio-oil increases [231]. Table 7 demonstrates that adding a catalyst reduces the pH of the bio-oil produced from pyrolyzing Brunei rice husk, whereas using a higher temperature raises the pH of the bio-oil produced from the pyrolyzing cotton stalk, where the pH of ZSM-5 is 2.74 at 450 °C and 3.9 at 600 °C. The pyrolysis of Pinus halepensis using CaO and a mixture of CaO and MgO, when their pH value is greater than non-catalytic bio-oil and compared to CaO, also increases the pH of the bio-oil from 2.8 (non-catalytic) to 4.8 (CaO). The loading effect of the calcium oxide catalyst was observed, where increasing the loading effect increased the pH value of the bio-oil. These results show that using basic oxide may increase the pH of the bio-oil [158,232,233,234].
The temperature has also an impact on the pH value, and temperature increases can raise the pH value of bio-oil. When catalysts are used, the pH can be increased or decreased, but with the help of temperature, the pH can increase due to the breakdown of acid components into smaller components. Utilizing basic oxide can also raise the pH of bio-oil.

5.2.2. Water Content

Water content refers to how much water is present in bio-oil; the lower the water content, the better the bio-oil is for use as a fuel for transportation. Table 8 illustrates that as the temperature rises, the water content of bio-oil produced by non-catalytic pyrolysis remains constant at 20–21% [231].
As per Table 8, the trend in water content is the same: using a catalyst results in more water being present in the bio-oil than not using one. This is because using a catalyst improves bio-oil by eliminating oxygen from oxygenating compounds in the form of water, carbon dioxide, and carbon monoxide [158,233,235]. It can be concluded that using catalysts can increase the water content of bio-oil due to the breakdown of oxygenated components into water and carbon dioxide.

5.2.3. High Heating Value (HHV)

The quantity of heat produced by the full combustion of a fuel is its heating value. The high heating value is the heat released from the fuel combustion with the original and generated water in a condensed state [236]. The higher the heating value, the more energy is formed and the more suitable it is to be used as a transportation fuel.
Bio-oil produced from the non-catalytic pyrolysis of pine wood chips is shown in Table 9, where an escalation in temperature slightly raises the HHV of the bio-oil. The HHV increased from 18.6 to 19.7 MJ/kg with the increase in temperature from 425 to 500 °C, indicating that the temperature has a minor role while catalyst and loading effects have a major role in increasing the high heating value [231].
The type of catalyst has an impact on the HHV values of biomass in the pyrolysis of Brunei rice husk. About 30% of the HHV value was increased with the change in catalyst type. The highest HHV was obtained at 17.17 MJ/kg for the BRHA catalyst. Moreover, an increase in the loading of the CaO catalyst increased the HHV value by 30% compared to the non-catalytic process, indicating the impact of catalyst loading on the increase in the HHV of the biomass in the pyrolysis process [158,234].

5.2.4. Elemental Analysis

The elemental or ultimate analysis is the measurement of the carbon-hydrogen-nitrogen-sulfur-oxygen-ash contained in the bio-oil. It can be used as an alternative method to determine the HHV of the bio-oil using the following formula, for example [237]:
H H V = 0.3491 C + 1.1783 H + 0.1005 S 0.1034 O 0.0151 N 0.0211 A
The bio-oil quality is better when it has a greater H/C and a lower O/C ratio.
From Table S3, as temperature increases, the carbon (C)% increases while hydrogen (H)%, and nitrogen (N)% decrease at 500 °C, then increases at 600 °C while oxygen (O)% increases, vice versa for the pyrolysis of raw pine bark in a non-catalytic process [238].
As shown in Table 10, the elemental analysis of the bio-oil manufactured by non-catalytic and catalytic pyrolysis with the increase in temperature was obtained, where utilizing a catalyst results in higher C% and N% in the catalytic process than without one at 350 °C, while for O%, it is further decreased in the catalytic process. While the temperature is augmented to 550 °C, C% is increasing and N% and O% are decreasing. Meanwhile, H% is fluctuating between 350 °C and 550 °C [239].
Hence, from Table 10, it can be concluded that increases in temperature with the help of a catalyst can increase the C% and decrease the O%, in which this change is due to the reactions that occur within the catalyst, such as dehydration and decarboxylation.

5.2.5. Stability of Bio-Oil

Biofuels are not as chemically or thermally stable as fossil fuels since they include a lot of oxygenated compounds. According to theory, the chemical interactions between components like ketones and aldehydes cause molecules with a high molecular weight to form, increasing the viscosity of bio-oil when temperatures are higher than 40 °C or when it is being stored for an extended period of time [240]. As a result, it is anticipated that bio-oil with a reduced carbonyl group concentration is more stable. Increased heavy-molecule-breaking processes and the removal of oxygenated chemicals result from the use of catalysts in the transportation and storage of bio-oil, producing bio-oil with a high degree of stability [241]. ZnO was used as a catalyst in a study by Nokkosmaki et al. [242] to pyrolyze pine sawdust. The stability test, which lasted 24 h at 80 °C, showed that the addition of ZnO changed the viscosity. The 55% change in viscosity was significantly smaller than the 129% change in non-catalytic bio-oil. Mante and Agblevor [243] examined the stability of the hybrid poplar wood-derived catalytic bio-oils (low and high bio-syncrude oils). A gravity oven was used for stability and aging evaluations, which lasted for 24 h at 90 °C. Moreover, the viscosities of the bio-oil samples stored at 40 °C for more than 10 months were analyzed. According to research, low bio-syncrude oil has a viscosity difference of 5%, whereas high bio-syncrude oil has a difference of 27.9%. It was discovered that catalytic bio-oils have excellent thermal stability and could be stored at room temperature for more than 10 months without viscosity appreciably rising.
When carbon dioxide (CO2) and water undergo catalytic pyrolysis, deoxygenation processes, such as dehydration, decarbonylation, and decarboxylation, release oxygen as a byproduct, regulating yield and quality [28]. By turning the oxygenated compounds into aromatic hydrocarbons, bio-oil produced by the pyrolysis of biomass is suited for upgraded bio-oil qualities. The bio-oils from catalytic pyrolysis are related to superior quality and greater output compared to non-catalytic pyrolysis due to the considerably larger proportions of aromatic hydrocarbons and pH. The quality of produced bio-oil often increases when a simple thermal process is switched to catalytic pyrolysis because of subsequent catalytic upgrading reactions.
To assess the economic viability of the pyrolysis process, a full grasp of cost concepts and their analysis is necessary. The outcomes of this investigation can help direct and influence future research and developments to boost efficacy and cost-effectiveness. We may make wise choices regarding the potential for commercialization and future expansion of pyrolysis as a promising alternative energy source by taking into account its technological, financial, and environmental ramifications.
Though pyrolysis has a lot of promise, commercialization is still difficult to achieve, and it is necessary to thoroughly comprehend its dynamics in order to scale up the process. Although difficult, time consuming, and expensive, waste biomass pyrolysis is still a viable option. If these challenges can be resolved, using machine learning as a potential tool for assisting and expediting pyrolysis research would make biomass pyrolysis an attractive choice.

6. Machine Learning and Techno-Economic Analysis of Catalytic Pyrolysis

6.1. Machine Learning

Computer programs may predict outcomes more correctly thanks to machine learning, a type of artificial intelligence (AI) that does not require any explicit training. In order to forecast new output values, machine learning algorithms use previous data as input. ML-based models just need an experimental data set and do not require a great deal of process understanding and information. Most often, artificial neural networks (ANNs) have been employed in the gasification and pyrolysis sectors [244]. A variety of machine learning algorithms are crucial in the data science sector. These machine learning methods are crucial for creating predictive models and performing categorization and prediction. For performing prediction and classification in both supervised and unsupervised contexts, these ML methods are the most beneficial. The most popular artificial intelligence algorithms are listed below [245]:
Artificial neural network;
Linear regression;
Decision trees;
Logistic regression;
Gaussian mixture model;
Naive Bayes;
K-means clustering;
Anomaly detection;
Principal component analysis;
K-nearest neighbors algorithm;
Support vector machines
Pyrolysis and gasification can be precisely modeled by machine learning models. To increase the production of renewable energy and meet national priorities like energy security and environmental sustainability, one of the numerous processes with optimization potential that may effectively use ML approaches is the thermal conversion of biomass [246]. More heed is being paid to the engineering applications that can benefit from process optimization by using the new knowledge that machine learning offers for managing the yields of pyrolytic products via the algorithm [247].
ANNs are a specific class of machine learning algorithms that mimic the human brain and function on the same fundamental principles as the neurons in our nervous system. The majority of recent research on the use of ML in pyrolysis is on traditional pyrolysis in fixed-bed reactors. The use of ML must be broadened, nevertheless, to include processes like catalytic pyrolysis and other types of reactors [248]. ML methods can forecast the yield and oxygen content of bio-oil depending on the properties of the biomass and the pyrolysis process [249], and the models based on artificial neural networks are also created to forecast biomass pyrolysis kinetics [250].
Four models, namely ANN-1 for proximal analysis; ANN-2 for ultimate analysis; ANN-3 for combined proximate and ultimate analysis; and ANN-4 for combined proximate, ultimate, and biochemical analysis, were employed by Balsora et al. [250]. With regression coefficient values of R2 ~ 0.99, relative root-mean-squared error ˂10.0%, and mean absolute error ˂0.071, it was discovered that the ANN-3 and ANN-4 models were highly useful in appropriately predicting the kinetic variables. The kinetic parameters discovered by ANN-3 and ANN-4 are in good agreement with the kinetics established using the analytical approach, which can be inferred from the discussion above.
ANN prediction was also shown to be superior to optimized response surface methodology for producing bio-oil from the pyrolysis of Mangifera indica wood sawdust. Based on average absolute relative deviation percentage = 0.25945, mean squared error = 0.00001, and R2= 0.99711, the ANN-FFBP [3-9-1] model demonstrated good agreement with experimental bio-oil yield, and the MIW bio-oil has a flash point of 70 °C, a kinematic viscosity of 3.97 cSt, a high heat capacity of 25.10 MJ/kg, a pH value of 2.78, a mass density of 1058.7 kg/m3, and a high heat index of 25.10 MJ/kg [251]. This finding implies that to help plant operators make judgments on process improvement, machine learning techniques might be applied more often.

6.2. Techno-Economic Analysis

Pyrolysis, gasification, and biochemical processes were subjected to techno-economic evaluations in order to assess their relevance in the conversion of biomass to liquid fuels [252]. The price of feedstock, the cost of transportation, the cost of utilities, and the cost of waste disposal systems are all variables that affect the cost of producing bio-oils from biomass by catalytic pyrolysis. Other constant costs include labor costs for installation and depreciation, labor costs for maintenance, and investment costs. The commercialization of bio-oil is dependent on plant capacity, feedstock availability, and government subsidies [253,254]. Depending on the sources, manufacturing processes, collection, processing, and transportation distance, feedstock expenses can range from USD 50 to USD 97 per ton, or between 30 and 54% of the cost of generating liquid fuels [255]. An investigation of the profitability of both slow and fast pyrolysis of common wood discovered that fast pyrolysis was economical as a result of increased production of bio-oil, which is highly valuable in the production of energy [256]. To assess the techno-economic performance, such as production costs and profitability, machine learning models may be trained using data on process parameters, feedstock characteristics, and other pertinent factors.
Cost is a crucial component in determining if a pyrolysis process is economically feasible, and both capital and operational expenses must be examined [257]. Operating expenses may be divided into fixed and variable costs. Utility bills, payroll taxes, rent or mortgage payments, and insurance are a few examples of fixed operating expenditures. A variable operating cost is the cost of supplies like byproducts, feedstock, and raw materials [258,259]. To assess the economic viability of the pyrolysis process, it is crucial to have a solid grasp of cost concepts and their analysis.
Rice straw collection is far less expensive than energy crop production [260]. About 17.7 wt% of fuel is produced by the mild catalytic pyrolysis of woody biomass, and 39% of the energy is converted to fuel. While the pyrolysis liquid needs 2.7 wt% hydrogens to be deoxygenated, it also needs 6.4 wt% hydrogens to be saturated with aromatic rings. For a plant that can process 2000 metric tons per day, the total project expenditure is USD 457 million, with operational expenses of USD 142 million per year. On the basis of a 10% internal rate of return, a minimum fuel selling price (MFSP) of USD 3.69 per gallon is predicted [261]. The techno-economic study of the catalytic pyrolysis of olive mill wastewater sludge using two distinct cooling techniques was conducted by Khan et al. [262], in which a vapor compression refrigeration machine (type-1) and an absorption refrigeration machine (type-2) were used. The production cost of bio-oil was achieved in the range of 2.16 EUR/GGE to 6.19 EUR/GGE for type-1 and 1.78 EUR/GGE and 5.01 EUR/GGE for type-2, where the cost parameters are adjusted within an industrially applicable range.
Fast pyrolysis had the largest yields and lowest production costs, the exact reverse of what happened with biological methods. From each liquid fuel plant, product values ranging from USD 0.53 to USD 1.45 per liter (USD 2.0 to USD 5.5 per gallon) were anticipated. For a facility processing 2000 tons of feedstock per day, product values were calculated using a maize stover costing USD 75 per ton and plant starting costs [252]. Fast pyrolysis analyses showed that power and biomass prices account for roughly 75% of production expenses. Selling the char generated by pyrolysis can lower the cost by around 18%. Production costs will rise from 6.8 GBP/GJ to 8.3 GBP/GJ with an 800-ton/day pyrolysis facility [263]. In the techno-economic analysis of catalytic pyrolysis for bio-oil production using ZSM-5, around 615 USD/ton was spent on making pyrolysis oil with 18% oxygen. When bio-oil yields rose by 20% (1000 tons/day), production costs dropped by up to 43 USD/ton. Pricing for biomass range from USD 57 to USD 114 per ton, and production expenses increased to as much as USD 841 per ton [264]. The economics of bio-oil production varies depending on the size, source, biomass, and catalyst expenses. As a result, the key to catalytic pyrolysis is the employment of affordable catalysts and close biomass sources [265].

7. Conclusions and Recommendations

7.1. Conclusions

In light of all the research results discussed in this review, it is noteworthy that the distinctive method for energy production from biomass is the catalytic pyrolysis process. This method can effectively transform waste materials into fuels and bioenergy with the support of catalysts. In the catalytic pyrolysis process, the quality of bio-oils is significantly increased by the formation of stable chemical compounds (phenolic, alcoholic, and aromatic compounds) and a reduction in unstable chemical compounds (acidic, oxygenated, and nitrogenous compounds) compared to the non-catalytic pyrolysis process. In the catalysts, zeolites are used mostly as a hydrocracking catalyst to increase the aromatic compounds in the bio-oil with higher thermal stability and better resistance to nitrogen and sulfur compounds. Among dolomite and sand as natural catalysts, the production of bio-oil was decreased in the catalytic pyrolysis process as the dolomite has the capability to break down heavy components. The activated carbon contained a huge number of active functional groups, which effectively increased the catalytic activity. The phenol components were significantly increased for breakdown into smaller components using activated carbon as catalysts. The mixture of catalysts is also effective for desirable outcomes in bio-oils. The quantity and purity of bio-oils generated by catalytic pyrolysis are significantly influenced by the structure of the biomass, the biomass-to-catalyst ratio, the operating temperature, and other process factors.
Catalytic fast pyrolysis is one of the most advantageous and affordable ways to extract upgraded bio-oil products from biomass feedstock. In the commercialization of bio-oils, the production costs are the key consideration in comparison to other biofuels, as the costs depend on the pretreatment, upgrading, and recycling methods. Techno-economic analysis, combined with the machine learning process, can effectively reduce production costs by enhancing the quality of the bio-oils. This process can also help direct the development of catalytic pyrolysis procedures that are both environmentally friendly and profitable. Finally, the catalytic pyrolysis process is one of the best methods to convert biomass into upgraded bioenergy, and techno-economic analysis with machine learning can enhance the viability of the process.

7.2. Recommendations for Future Works

The results of this study contribute significantly to the field of renewable energy science. The concept presented here may be applied to the pyrolysis process design and development of other biomass samples for energy generation. According to the experiments and findings, the study has met the expectations within the framework, and it can now be advanced for the future as follows:
  • The catalytic effect can be observed with different biomass-to-catalyst ratios with in situ and ex situ arrangements in the future;
  • Inexpensive materials can be studied as catalysts for catalytic pyrolysis;
  • The experiment can be optimized with higher accuracy by changing the operating parameters like heating rate, nitrogen flow rate, etc.;
  • The use of regenerated catalysts can be continued several times until satisfactory results are achieved;
  • Different mixture ratios can be applied with activated carbon and other low-cost metal oxides to obtain upgraded bio-oils
  • The syngas can be analyzed for further applications in the fuel cell to obtain clean energy;
  • The bio-oils produced from catalytic pyrolysis can be used as fuel in real-world applications to achieve practical performance;
  • A gap between lab-scale studies and the real industrial situation can be thoroughly discussed;
  • Standardization of raw materials and process designs for the catalytic pyrolysis process should be performed with optimum techno-economic solutions;
  • In order to expedite catalytic pyrolysis research, extensive machine learning processes can be explored due to their capacity to learn nonlinear input and output correlations;
  • Future models for process optimization and control ought to strike a balance between determining load and precision. However, ML is not a universal fix for all issues in pyrolysis research; nevertheless, it may work in conjunction with current technologies and provide a great tool for pyrolysis process monitoring, control, and product yield optimization.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/1996-1073/16/14/5547/s1, Table S1: Conditions of operation and product yields of several pyrolysis processes; Table S2: Advantages and disadvantages of various pyrolysis technologies; Table S3: Elements of the bio-oil produced by the pyrolysis of raw pine bark.

Author Contributions

Conceptualization M.S.R., S.A. and A.K.A.; writing—original draft preparation, M.S.R., S.A., A.K., K.Z.B., M.M.K., H.R. and M.S.I.; writing—review and editing, Z.B.I., M.S.A.B., A.K.A. and M.S.I.; supervision, Z.B.I., M.S.A.B., A.K.A. and K.K.; funding acquisition, K.K. and M.S.I. All authors have read and agreed to the published version of the manuscript.

Funding

The research was completed as part of Project No. AP09261208 supported by the Ministry of Science and Higher Education, Kazakhstan.

Data Availability Statement

Not applicable.

Acknowledgments

The authors highly acknowledge L.N. Gumilyov Eurasian National University, Kazakhstan, Universiti Brunei Darussalam, Brunei, and Bangladesh University of Engineering and Technology for facilitating this work.

Conflicts of Interest

The authors declare no conflict of interest.

Glossary

Nomenclature
AlAluminum
Al2O3Aluminum oxides
CaCalcium
CaMg(CO3)2Dolomite
CaOCalcium oxide
CeO2Cerium dioxide
CO2Carbon dioxide
COxCarbon oxides
COCarbon monoxide
(C6H10O5)nCellulose
(C5H8O4)nHemicellulose
CH4Methane
CrChromium
FeIron
Fe2O3Ferric oxide
H2Hydrogen
H2OWater
HClHydrogen chloride
KPotassium
KClPotassium chloride
K2Cl2Potassium chloride
K3PO4Tripotassium phosphate
MgMagnesium
MgOMagnesium oxide
MnManganese
MnO2Manganese dioxide
N2Nitrogen
NaSodium
NaClSodium chloride
PPhosphorus
R2Regression coefficient
SiSilicon
SiCSilicon carbide
SiO2Silicon dioxide
TiO2Titanium dioxide
ZSM-5Zeolite Socony Mobil–5
ZnZinc
ZnOZinc oxide
Abbreviations
3DThree dimensional
ACActivated carbon
AIArtificial intelligence
ANNArtificial neural network
Bentonite Aluminum phyllosilicate
BRHABrunei rice husk ash
HHVHigh heating value
MLMachine learning
TEATechno-economic analysis

References

  1. Mohabeer, C.; Guilhaume, N.; Laurenti, D.; Schuurman, Y. Microwave-Assisted Pyrolysis of Biomass with and without Use of Catalyst in a Fluidised Bed Reactor: A Review. Energies 2022, 15, 3258. [Google Scholar] [CrossRef]
  2. Alazaiza, M.Y.D.; Albahnasawi, A.; Al Maskari, T.; Abujazar, M.S.S.; Bashir, M.J.K.; Nassani, D.E.; Abu Amr, S.S. Biofuel Production Using Cultivated Algae: Technologies, Economics, and Its Environmental Impacts. Energies 2023, 16, 1316. [Google Scholar] [CrossRef]
  3. Afroze, S.; Reza, M.S.; Amin, M.R.; Taweekun, J.; Azad, A.K. Progress in Nanomaterials Fabrication and Their Prospects in Artificial Intelligence towards Solid Oxide Fuel Cells: A Review. Int. J. Hydrogen Energy 2022, in press. [Google Scholar] [CrossRef]
  4. Ferreira, A.P.R.A.; Oliveira, R.C.P.; Mateus, M.M.; Santos, D.M.F.; Ferreira, A.P.R.A.; Oliveira, R.C.P.; Mateus, M.M.; Santos, D.M.F. A Review of the Use of Electrolytic Cells for Energy and Environmental Applications. Energies 2023, 16, 1593. [Google Scholar] [CrossRef]
  5. Reza, M.S.; Afroze, S.; Kuterbekov, K.; Kabyshev, A.; Bekmyrza, K.Z.; Haque, M.N.; Islam, S.N.; Hossain, M.A.; Hassan, M.; Roy, H.; et al. Advanced Applications of Carbonaceous Materials in Sustainable Water Treatment, Energy Storage, and CO2 Capture: A Comprehensive Review. Sustainability 2023, 15, 8815. [Google Scholar] [CrossRef]
  6. Perišić, M.; Barceló, E.; Dimic-Misic, K.; Imani, M.; Spasojević Brkić, V. The Role of Bioeconomy in the Future Energy Scenario: A State-of-the-Art Review. Sustainability 2022, 14, 560. [Google Scholar] [CrossRef]
  7. Raza, M.; Inayat, A.; Ahmed, A.; Jamil, F.; Ghenai, C.; Naqvi, S.R.; Shanableh, A.; Ayoub, M.; Waris, A.; Park, Y.K. Progress of the Pyrolyzer Reactors and Advanced Technologies for Biomass Pyrolysis Processing. Sustainability 2021, 13, 11061. [Google Scholar] [CrossRef]
  8. Syazaidah, I.; Abu Bakar, M.S.; Reza, M.S.; Azad, A.K. Ex-Situ Catalytic Pyrolysis of Chicken Litter for Bio-Oil Production: Experiment and Characterization. J. Environ. Manag. 2021, 297, 113407. [Google Scholar] [CrossRef]
  9. Dickerson, T.; Soria, J. Catalytic Fast Pyrolysis: A Review. Energies 2013, 6, 514–538. [Google Scholar] [CrossRef]
  10. van Meerbeek, K.; Appels, L.; Dewil, R.; Calmeyn, A.; Lemmens, P.; Muys, B.; Hermy, M.; Van Meerbeek, K.; Appels, L.; Dewil, R.; et al. Biomass of Invasive Plant Species as a Potential Feedstock for Bioenergy Production. Biofuels Bioprod. Biorefining 2015, 9, 273–282. [Google Scholar] [CrossRef]
  11. Chen, Y. Biomass to Fuels: Thermo-Chemical or Bio-Chemical Conversion? Ferment. Technol. 2012, 1, 7972. [Google Scholar] [CrossRef]
  12. Halder, P.; Azad, A.K. Recent Trends and Challenges of Algal Biofuel Conversion Technologies. In Advanced Biofuels: Applications, Technologies and Environmental Sustainability; Elsevier: Amsterdam, The Netherlands, 2019; pp. 169–179. ISBN 9780081027912. [Google Scholar]
  13. Panwar, N.L.; Kothari, R.; Tyagi, V.V. Thermo Chemical Conversion of Biomass—Eco Friendly Energy Routes. Renew. Sustain. Energy Rev. 2012, 16, 1801–1816. [Google Scholar] [CrossRef]
  14. Zaman, C.Z.; Pal, K.; Yehye, W.A.; Sagadevan, S.; Shah, S.T.; Adebisi, G.A.; Marliana, E.; Rafique, R.F.; Johan, R.B. Pyrolysis: A Sustainable Way to Generate Energy from Waste. In Pyrolysis; InTech: London, UK, 2017; pp. 1–35. ISBN 978-953-51-3312-4. [Google Scholar]
  15. Chen, W.H.; Lin, B.J. Characteristics of Products from the Pyrolysis of Oil Palm Fiber and Its Pellets in Nitrogen and Carbon Dioxide Atmospheres. Energy 2016, 94, 569–578. [Google Scholar] [CrossRef]
  16. Wainaina, S.; Awasthi, M.K.; Sarsaiya, S.; Chen, H.; Singh, E.; Kumar, A.; Ravindran, B.; Awasthi, S.K.; Liu, T.; Duan, Y.; et al. Resource Recovery and Circular Economy from Organic Solid Waste Using Aerobic and Anaerobic Digestion Technologies. Bioresour. Technol. 2020, 301, 122778. [Google Scholar] [CrossRef]
  17. Wang, C.; Huang, R.; Sun, R.; Yang, J.; Sillanpää, M. A Review on Persulfates Activation by Functional Biochar for Organic Contaminants Removal: Synthesis, Characterizations, Radical Determination, and Mechanism. J. Environ. Chem. Eng. 2021, 9, 106267. [Google Scholar] [CrossRef]
  18. Reza, M.S.; Hasan, A.B.M.K.; Ahmed, A.S.; Afroze, S.; Bakar, M.S.A.; Islam, S.N.; Azad, A.K. COVID-19 Prevention: Role of Activated Carbon. J. Eng. Technol. Sci. 2021, 53, 210404. [Google Scholar] [CrossRef]
  19. Elliott, D.C. Historical Developments in Hydroprocessing Bio-Oils. Energy Fuels 2007, 21, 1792–1815. [Google Scholar] [CrossRef]
  20. Afroze, S.; Torino, N.; Reza, M.S.; Radenahmad, N.; Cheok, Q.; Henry, P.F.; Azad, A.K. Structure-Conductivity Relationship of PrBaMnMoO6-δ through in-Situ Measurements: A Neutron Diffraction Study. Ceram. Int. 2021, 47, 541–546. [Google Scholar] [CrossRef]
  21. Balat, M. An Overview of the Properties and Applications of Biomass Pyrolysis Oils. Energy Sources Part A Recover. Util. Environ. Eff. 2011, 33, 674–689. [Google Scholar] [CrossRef]
  22. Kan, T.; Strezov, V.; Evans, T.; He, J.; Kumar, R.; Lu, Q. Catalytic Pyrolysis of Lignocellulosic Biomass: A Review of Variations in Process Factors and System Structure. Renew. Sustain. Energy Rev. 2020, 134, 110305. [Google Scholar] [CrossRef]
  23. Li, P.; Niu, B.; Pan, H.; Zhang, Y.; Long, D. Production of Hydrocarbon-Rich Bio-Oil from Catalytic Pyrolysis of Waste Cooking Oil over Nickel Monoxide Loaded Corn Cob-Derived Activated Carbon. J. Clean. Prod. 2023, 384, 135653. [Google Scholar] [CrossRef]
  24. Chen, X.; Che, Q.; Li, S.; Liu, Z.; Yang, H.; Chen, Y.; Wang, X.; Shao, J.; Chen, H. Recent Developments in Lignocellulosic Biomass Catalytic Fast Pyrolysis: Strategies for the Optimization of Bio-Oil Quality and Yield. Fuel Process. Technol. 2019, 196, 106180. [Google Scholar] [CrossRef]
  25. Duan, Y.; Pandey, A.; Zhang, Z.; Awasthi, M.K.; Bhatia, S.K.; Taherzadeh, M.J. Organic Solid Waste Biorefinery: Sustainable Strategy for Emerging Circular Bioeconomy in China. Ind. Crops Prod. 2020, 153, 112568. [Google Scholar] [CrossRef]
  26. Reza, M.S.; Afroze, S.; Kuterbekov, K.; Kabyshev, A.; Bekmyrza, K.Z.; Taweekun, J.; Ja’afar, F.; Bakar, M.S.A.; Azad, A.K.; Roy, H.; et al. Ex Situ Catalytic Pyrolysis of Invasive Pennisetum Purpureum Grass with Activated Carbon for Upgrading Bio-Oil. Sustainability 2023, 15, 7628. [Google Scholar] [CrossRef]
  27. Nishu; Liu, R.; Rahman, M.M.; Sarker, M.; Chai, M.; Li, C.; Cai, J. A Review on the Catalytic Pyrolysis of Biomass for the Bio-Oil Production with ZSM-5: Focus on Structure. Fuel Process. Technol. 2020, 199, 106301. [Google Scholar] [CrossRef]
  28. Jain, A.; Sarsaiya, S.; Kumar Awasthi, M.; Singh, R.; Rajput, R.; Mishra, U.C.; Chen, J.; Shi, J. Bioenergy and Bio-Products from Bio-Waste and Its Associated Modern Circular Economy: Current Research Trends, Challenges, and Future Outlooks. Fuel 2022, 307, 121859. [Google Scholar] [CrossRef]
  29. Elmaz, F.; Yücel, Ö.; Mutlu, A.Y. Predictive Modeling of Biomass Gasification with Machine Learning-Based Regression Methods. Energy 2020, 191, 116541. [Google Scholar] [CrossRef]
  30. Jawaid, M.; Paridah, M.T.; Saba, N. Introduction to Biomass and Its Composites. In Lignocellulosic Fibre and Biomass-Based Composite Materials; Woodhead Publishing: Sawston, UK, 2017; pp. 1–11. [Google Scholar] [CrossRef]
  31. Pampillón-González, L.; Canepa, J.R.L. Biomass as an Alternative for Gas Production. In Advances in Natural Gas Emerging Technologies; InTech: London, UK, 2017. [Google Scholar]
  32. Tumuluru, J.S.; Wright, C.T.; Boardman, R.D.; Yancey, N.A.; Sokhansanj, S. A Review on Biomass Classification and Composition, Co-Firing Issues and Pretreatment Methods. In Proceedings of the 2011 ASABE Annual International Meeting, Louisville, KY, USA, 7–10 August 2011; American Society of Agricultural and Biological Engineers: St. Joseph, MI, USA, 2011. [Google Scholar]
  33. Mansor, A.M.; Lim, J.S.; Ani, F.N.; Hashim, H.; Ho, W.S. Characteristics of Cellulose, Hemicellulose and Lignin of MD2 Pineapple Biomass. Chem. Eng. Trans. 2019, 72, 79–84. [Google Scholar] [CrossRef]
  34. Thakur, V.K.; Thakur, M.K. Recent Advances in Green Hydrogels from Lignin: A Review. Int. J. Biol. Macromol. 2015, 72, 834–847. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, K.; Kim, K.H.; Brown, R.C. Catalytic Pyrolysis of Individual Components of Lignocellulosic Biomass. Green Chem. 2014, 16, 727–735. [Google Scholar] [CrossRef]
  36. Zhang, S.; Sun, J.; Zhang, X.; Xin, J.; Miao, Q.; Wang, J. Ionic Liquid-Based Green Processes for Energy Production. Chem. Soc. Rev. 2014, 43, 7838–7869. [Google Scholar] [CrossRef] [PubMed]
  37. Preparation and Characterization of Polyvinyl Alcohol-Based Composite Reinforced with Nanocellulose and Nanosilica|Ching|BioResources. Available online: https://ojs.cnr.ncsu.edu/index.php/BioRes/article/view/6992 (accessed on 11 February 2021).
  38. Sampath, U.G.T.M.; Ching, Y.C.; Chuah, C.H.; Singh, R.; Lin, P.C. Preparation and Characterization of Nanocellulose Reinforced Semi-Interpenetrating Polymer Network of Chitosan Hydrogel. Cellulose 2017, 24, 2215–2228. [Google Scholar] [CrossRef]
  39. Guedes, R.E.; Luna, A.S.; Torres, A.R. Operating Parameters for Bio-Oil Production in Biomass Pyrolysis: A Review. J. Anal. Appl. Pyrolysis 2018, 129, 134–149. [Google Scholar] [CrossRef]
  40. Huang, W.; Gong, F.; Fan, M.; Zhai, Q.; Hong, C.; Li, Q. Production of Light Olefins by Catalytic Conversion of Lignocellulosic Biomass with HZSM-5 Zeolite Impregnated with 6 wt.% Lanthanum. Bioresour. Technol. 2012, 121, 248–255. [Google Scholar] [CrossRef]
  41. Peters, B. Prediction of Pyrolysis of Pistachio Shells Based on Its Components Hemicellulose, Cellulose and Lignin. Fuel Process. Technol. 2011, 92, 1993–1998. [Google Scholar] [CrossRef]
  42. Mihalcik, D.J.; Mullen, C.A.; Boateng, A.A. Screening Acidic Zeolites for Catalytic Fast Pyrolysis of Biomass and Its Components. J. Anal. Appl. Pyrolysis 2011, 92, 224–232. [Google Scholar] [CrossRef]
  43. Wendisch, V.F.; Kim, Y.; Lee, J.H. Chemicals from Lignin: Recent Depolymerization Techniques and Upgrading Extended Pathways. Curr. Opin. Green Sustain. Chem. 2018, 14, 33–39. [Google Scholar] [CrossRef]
  44. Yang, J.; Ching, Y.C.; Chuah, C.H. Applications of Lignocellulosic Fibers and Lignin in Bioplastics: A Review. Polymers 2019, 11, 751. [Google Scholar] [CrossRef]
  45. Rahimi, A.; Ulbrich, A.; Coon, J.J.; Stahl, S.S. Formic-Acid-Induced Depolymerization of Oxidized Lignin to Aromatics. Nature 2014, 515, 249–252. [Google Scholar] [CrossRef]
  46. Cagnon, B.; Py, X.; Guillot, A.; Stoeckli, F.; Chambat, G. Contributions of Hemicellulose, Cellulose and Lignin to the Mass and the Porous Properties of Chars and Steam Activated Carbons from Various Lignocellulosic Precursors. Bioresour. Technol. 2009, 100, 292–298. [Google Scholar] [CrossRef]
  47. Yang, H.; Yan, R.; Chen, H.; Lee, D.H.; Zheng, C. Characteristics of Hemicellulose, Cellulose and Lignin Pyrolysis. Fuel 2007, 86, 1781–1788. [Google Scholar] [CrossRef]
  48. Priharto, N.; Ronsse, F.; Yildiz, G.; Heeres, H.J.; Deuss, P.J.; Prins, W. Fast Pyrolysis with Fractional Condensation of Lignin-Rich Digested Stillage from Second-Generation Bioethanol Production. J. Anal. Appl. Pyrolysis 2020, 145, 104756. [Google Scholar] [CrossRef]
  49. Dhyani, V.; Bhaskar, T. A Comprehensive Review on the Pyrolysis of Lignocellulosic Biomass. Renew. Energy 2018, 129, 695–716. [Google Scholar] [CrossRef]
  50. Williams, C.L.; Westover, T.L.; Emerson, R.M.; Tumuluru, J.S.; Li, C.; Westover, T.L.; Emerson, R.M.; Tumuluru, J.S.; Li, C. Sources of Biomass Feedstock Variability and the Potential Impact on Biofuels Production. Bioenergy Res. 2016, 9, 1–14. [Google Scholar] [CrossRef]
  51. Pecha, B.; Garcia-Perez, M. Pyrolysis of Lignocellulosic Biomass. In Bioenergy; Elsevier: London, UK, 2015; pp. 413–442. [Google Scholar]
  52. Mohan, D.; Pittman, C.U.; Steele, P.H. Pyrolysis of Wood/biomass for Bio-Oil: A Critical Review. Energy Fuels 2006, 20, 848–889. [Google Scholar] [CrossRef]
  53. Antal, M.J. The Art, Science, and Technology of Charcoal Production. Ind. Eng. Chem. Res. 2003, 42, 1619–1640. [Google Scholar] [CrossRef]
  54. Pattiya, A. Catalytic Pyrolysis. In Direct Thermochemical Liquefaction for Energy Applications; Elsevier Ltd.: Duxford, UK, 2018; pp. 29–64. ISBN 9780081010259. [Google Scholar]
  55. Patel, M.; Zhang, X.; Kumar, A. Techno-Economic and Life Cycle Assessment on Lignocellulosic Biomass Thermochemical Conversion Technologies: A Review. Renew. Sustain. Energy Rev. 2016, 53, 1486–1499. [Google Scholar] [CrossRef]
  56. Bridgwater, T. Biomass for Energy. J. Sci. Food Agric. 2006, 86, 1755–1768. [Google Scholar] [CrossRef]
  57. Akhtar, A.; Krepl, V.; Ivanova, T. A Combined Overview of Combustion, Pyrolysis, and Gasification of Biomass. Energy Fuels 2018, 32, 7294–7318. [Google Scholar] [CrossRef]
  58. Lackner, M. Combustion. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2011. [Google Scholar]
  59. Dincer, I.; Zamfirescu, C. Conventional Power Generating Systems. In Advanced Power Generation Systems; Elsevier: Amsterdam, The Netherlands, 2014; pp. 199–310. [Google Scholar]
  60. Munawer, M.E. Human Health and Environmental Impacts of Coal Combustion and Post-Combustion Wastes. J. Sustain. Min. 2018, 17, 87–96. [Google Scholar] [CrossRef]
  61. Jerzak, W.; Kuźnia, M. Examination of Inorganic Gaseous Species and Condensed Phases during Coconut Husk Combustion Based on Thermodynamic Equilibrium Predictions. Renew. Energy 2021, 167, 497–507. [Google Scholar] [CrossRef]
  62. Radenahmad, N.; Tasfiah, A.; Saghir, M.; Taweekun, J.; Saifullah, M.; Bakar, A.; Reza, S.; Kalam, A. A Review on Biomass Derived Syngas for SOFC Based Combined Heat and Power Application. Renew. Sustain. Energy Rev. 2020, 119, 109560. [Google Scholar] [CrossRef]
  63. Breault, R.W. Gasification Processes Old and New: A Basic Review of the Major Technologies. Energies 2010, 3, 216–240. [Google Scholar] [CrossRef]
  64. Afroze, S.; Reza, M.S.; Cheok, Q.; Islam, S.N.; Abdalla, A.M.; Taweekun, J.; Azad, A.K.; Khalilpoor, N.; Issakhov, A. Advanced Applications of Fuel Cells during the COVID-19 Pandemic. Int. J. Chem. Eng. 2021, 2021, 1–9. [Google Scholar] [CrossRef]
  65. Costa, P.; Pinto, F.; André, R.N.; Marques, P. Integration of Gasification and Solid Oxide Fuel Cells (SOFCs) for Combined Heat and Power (CHP). Processes 2021, 9, 254. [Google Scholar] [CrossRef]
  66. Reza, M.S.; Taweekun, J.; Afroze, S.; Siddique, S.A.; Islam, M.S.; Wang, C.; Azad, A.K. Investigation of Thermochemical Properties and Pyrolysis of Barley Waste as a Source for Renewable Energy. Sustainability 2023, 15, 1643. [Google Scholar] [CrossRef]
  67. Jabar, J.M.; Jabar, J.M. Pyrolysis: A Convenient Route for Production of Eco-Friendly Fuels and Precursors for Chemical and Allied Industries. In Recent Perspectives in Pyrolysis Research; IntechOpen: London, UK, 2021; ISBN 978-1-83969-915-3. [Google Scholar]
  68. Husain, Z.; Ansari, K.B.; Chatake, V.S.; Urunkar, Y.; Pandit, A.B.; Joshi, J.B. Valorisation of Biomass Pellets to Renewable Fuel and Chemicals Using Pyrolysis: Characterisation of Pyrolysis Products and Its Application. Indian Chem. Eng. 2020, 62, 78–91. [Google Scholar] [CrossRef]
  69. AlNouss, A.; McKay, G.; Al-Ansari, T. Production of Syngas via Gasification Using Optimum Blends of Biomass. J. Clean. Prod. 2020, 242, 118499. [Google Scholar] [CrossRef]
  70. Bhatia, S.C. Biogas. In Advanced Renewable Energy Systems; Woodhead Publishing: Sawston, UK, 2014; pp. 426–472. ISBN 978-1-78242-269-3. [Google Scholar]
  71. Tripathi, M.; Sahu, J.N.; Ganesan, P. Effect of Process Parameters on Production of Biochar from Biomass Waste through Pyrolysis: A Review. Renew. Sustain. Energy Rev. 2016, 55, 467–481. [Google Scholar] [CrossRef]
  72. Roy, P.; Dias, G. Prospects for Pyrolysis Technologies in the Bioenergy Sector: A Review. Renew. Sustain. Energy Rev. 2017, 77, 59–69. [Google Scholar] [CrossRef]
  73. Pecha, M.B.; Garcia-Perez, M. Pyrolysis of Lignocellulosic Biomass: Oil, Char, and Gas. In Bioenergy; Elsevier: London, UK, 2020; pp. 581–619. [Google Scholar]
  74. Sarangi, P.K.; Nanda, S.; Mohanty, P. Recent Advancements in Biofuels and Bioenergy Utilization; Springer: Singapore, 2018; ISBN 9789811313073. [Google Scholar]
  75. Torri, I.D.V.; Paasikallio, V.; Faccini, C.S.; Huff, R.; Caramão, E.B.; Sacon, V.; Oasmaa, A.; Zini, C.A. Bio-Oil Production of Softwood and Hardwood Forest Industry Residues through Fast and Intermediate Pyrolysis and Its Chromatographic Characterization. Bioresour. Technol. 2016, 200, 680–690. [Google Scholar] [CrossRef] [PubMed]
  76. Acosta, R.; Tavera, C.; Gauthier-Maradei, P.; Nabarlatz, D. Production of Oil and Char by Intermediate Pyrolysis of Scrap Tyres: Influence on Yield and Product Characteristics. Int. J. Chem. React. Eng. 2015, 13, 189–200. [Google Scholar] [CrossRef]
  77. Bridgwater, A.V.; Meier, D.; Radlein, D. An Overview of Fast Pyrolysis of Biomass. Org. Geochem. 1999, 30, 1479–1493. [Google Scholar] [CrossRef]
  78. Funke, A.; Tomasi Morgano, M.; Dahmen, N.; Leibold, H. Experimental Comparison of Two Bench Scale Units for Fast and Intermediate Pyrolysis. J. Anal. Appl. Pyrolysis 2017, 124, 504–514. [Google Scholar] [CrossRef]
  79. Liang, J.; Shan, G.; Sun, Y. Catalytic Fast Pyrolysis of Lignocellulosic Biomass: Critical Role of Zeolite Catalysts. Renew. Sustain. Energy Rev. 2021, 139, 110707. [Google Scholar] [CrossRef]
  80. Li, L.; Rowbotham, J.S.; Christopher Greenwell, H.; Dyer, P.W. An Introduction to Pyrolysis and Catalytic Pyrolysis: Versatile Techniques for Biomass Conversion. In New and Future Developments in Catalysis: Catalytic Biomass Conversion; Elsevier B.V.: Amsterdam, Netherlands, 2013; pp. 173–208. ISBN 9780444538789. [Google Scholar]
  81. Hornung, A. Intermediate Pyrolysis of Biomass. In Biomass Combustion Science, Technology and Engineering; Woodhead Publishing Limited: Sawston, UK, 2013; pp. 172–186. ISBN 9780857091314. [Google Scholar]
  82. Garcia-Perez, M.; Garcia-Nunez, J.A.; Lewis, T.; Kruger, C.; Kantor, S. Methods for Producing Biochar and Advanced Bio-Fuels in Washington State; Washington State University: Pullman, WA, USA, 2010; ISBN 3604076900. [Google Scholar]
  83. Bridgwater, A.V.; Peacocke, G.V.C. Fast Pyrolysis Processes for Biomass. Renew. Sustain. Energy Rev. 2000, 4, 1–73. [Google Scholar] [CrossRef]
  84. Warnecke, R. Gasification of Biomass: Comparison of Fixed Bed and Fluidized Bed Gasifier. Biomass Bioenergy 2000, 18, 489–497. [Google Scholar] [CrossRef]
  85. Cornelissen, G.; Pandit, N.R.; Taylor, P.; Pandit, B.H.; Sparrevik, M.; Schmidt, H.P. Emissions and Char Quality of Flame-curtain “Kon Tiki” kilns for Farmer-Scale Charcoal/biochar Production. PLoS ONE 2016, 11, e0154617. [Google Scholar] [CrossRef]
  86. Campuzano, F.; Brown, R.C.; Martínez, J.D. Auger Reactors for Pyrolysis of Biomass and Wastes. Renew. Sustain. Energy Rev. 2019, 102, 372–409. [Google Scholar] [CrossRef]
  87. Li, J.; Dai, J.; Liu, G.; Zhang, H.; Gao, Z.; Fu, J.; He, Y.; Huang, Y. Biochar from Microwave Pyrolysis of Biomass: A Review. Biomass Bioenergy 2016, 94, 228–244. [Google Scholar] [CrossRef]
  88. Jahirul, M.I.; Rasul, M.G.; Chowdhury, A.A.; Ashwath, N. Biofuels Production through Biomass Pyrolysis—A Technological Review. Energies 2012, 5, 4952–5001. [Google Scholar] [CrossRef]
  89. Uddin, M.; Joardder, M.U.H.; Islam, M.N. Design and Construction of Fixed Bed Pyrolysis System and Plum Seed Pyrolysis for Bio-Oil Production. Int. J. Adv. Renew. Energy Res. 2012, 1, 405–409. [Google Scholar]
  90. Bamido, A. Design of A Fluidized Bed Reactor for Biomass Pyrolysis. Master’s Thesis, University of Cincinnati, Cincinnati, OH, USA, 2019. [Google Scholar]
  91. Roegiers, J. Heat and Mass Transfer Modelling of Auger Reactors. Master’s Thesis, Universiteit Gent, Ghent, Belgium, 2016. [Google Scholar]
  92. Garcia-Perez, M.; Adams, T.T.; Goodrum, J.W.; Geller, D.; Das, K.C. Production and Fuel Properties of Pine Chip Bio-Oil/biodiesel Blends. Energy Fuels 2007, 21, 2363–2372. [Google Scholar] [CrossRef]
  93. Verma, M.; Godbout, S.; Brar, S.K.; Solomatnikova, O.; Lemay, S.P.; Larouche, J.P. Biofuels Production from Biomass by Thermochemical Conversion Technologies. Int. J. Chem. Eng. 2012, 2012, 542426. [Google Scholar] [CrossRef]
  94. Lappas, A.; Heracleous, E. Production of Biofuels via Fischer-Tropsch Synthesis: Biomass-to-Liquids; Woodhead Publishing Limited: Cambridge, UK, 2011; ISBN 9780857090492. [Google Scholar]
  95. Morin, M.; Pécate, S.; Hémati, M.; Kara, Y. Pyrolysis of Biomass in a Batch Fluidized Bed Reactor: Effect of the Pyrolysis Conditions and the Nature of the Biomass on the Physicochemical Properties and the Reactivity of Char. J. Anal. Appl. Pyrolysis 2016, 122, 511–523. [Google Scholar] [CrossRef]
  96. Fernández, Y.; Menéndez, J.A. Influence of Feed Characteristics on the Microwave-Assisted Pyrolysis Used to Produce Syngas from Biomass Wastes. J. Anal. Appl. Pyrolysis 2011, 91, 316–322. [Google Scholar] [CrossRef]
  97. Lam, S.S.; Russell, A.D.; Chase, H.A. Microwave Pyrolysis, a Novel Process for Recycling Waste Automotive Engine Oil. Energy 2010, 35, 2985–2991. [Google Scholar] [CrossRef]
  98. Aho, A.; Tokarev, A.; Backman, P.; Kumar, N.; Eränen, K.; Hupa, M.; Holmbom, B.; Salmi, T.; Murzin, D.Y. Catalytic Pyrolysis of Pine Biomass over H-Beta Zeolite in a Dual-Fluidized Bed Reactor: Effect of Space Velocity on the Yield and Composition of Pyrolysis Products. Top. Catal. 2011, 54, 941–948. [Google Scholar] [CrossRef]
  99. Atutxa, A.; Aguado, R.; Gayubo, A.G.; Olazar, M.; Bilbao, J. Kinetic Description of the Catalytic Pyrolysis of Biomass in a Conical Spouted Bed Reactor. Energy Fuels 2005, 19, 765–774. [Google Scholar] [CrossRef]
  100. Lappas, A.A.; Samolada, M.C.; Iatridis, D.K.; Voutetakis, S.S.; Vasalos, I.A. Biomass Pyrolysis in a Circulating Fluid Bed Reactor for the Production of Fuels and Chemicals. Fuel 2002, 81, 2087–2095. [Google Scholar] [CrossRef]
  101. Wang, J.X.; Cao, J.P.; Zhao, X.Y.; Liu, S.N.; Ren, X.Y.; Zhao, M.; Cui, X.; Chen, Q.; Wei, X.Y. Enhancement of Light Aromatics from Catalytic Fast Pyrolysis of Cellulose over Bifunctional Hierarchical HZSM-5 Modified by Hydrogen Fluoride and Nickel/hydrogen Fluoride. Bioresour. Technol. 2019, 278, 116–123. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Yi, W.; Fu, P.; Li, Z.; Wang, N.; Tian, C. Numerical Simulation and Experiment on Catalytic Upgrading of Biomass Pyrolysis Vapors in V-Shaped Downer Reactors. Bioresour. Technol. 2019, 274, 207–214. [Google Scholar] [CrossRef] [PubMed]
  103. Naqvi, S.R.; Tariq, R.; Shahbaz, M.; Naqvi, M.; Aslam, M.; Khan, Z.; Mackey, H.; Mckay, G.; Al-Ansari, T. Recent Developments on Sewage Sludge Pyrolysis and Its Kinetics: Resources Recovery, Thermogravimetric Platforms, and Innovative Prospects. Comput. Chem. Eng. 2021, 150, 107325. [Google Scholar] [CrossRef]
  104. Xiu, S.; Shahbazi, A. Bio-Oil Production and Upgrading Research: A Review. Renew. Sustain. Energy Rev. 2012, 16, 4406–4414. [Google Scholar] [CrossRef]
  105. Tomczyk, A.; Sokołowska, Z.; Boguta, P. Biochar Physicochemical Properties: Pyrolysis Temperature and Feedstock Kind Effects. Rev. Environ. Sci. Biotechnol. 2020, 19, 191–215. [Google Scholar] [CrossRef]
  106. Pal, S.; Kumar, A.; Sharma, A.K.; Ghodke, P.K.; Pandey, S.; Patel, A. Recent Advances in Catalytic Pyrolysis of Municipal Plastic Waste for the Production of Hydrocarbon Fuels. Processes 2022, 10, 1497. [Google Scholar] [CrossRef]
  107. Elliott, D.C.; Hart, T.R.; Neuenschwander, G.G.; Rotness, L.J.; Olarte, M.V.; Zacher, A.H.; Solantausta, Y. Catalytic Hydroprocessing of Fast Pyrolysis Bio-Oil from Pine Sawdust. Energy Fuels 2012, 26, 3891–3896. [Google Scholar] [CrossRef]
  108. Czernik, S.; Bridgwater, A.V. Overview of Applications of Biomass Fast Pyrolysis Oil. Energy Fuels 2004, 18, 590–598. [Google Scholar] [CrossRef]
  109. Treedet, W.; Suntivarakorn, R. Design and Operation of a Low Cost Bio-Oil Fast Pyrolysis from Sugarcane Bagasse on Circulating Fluidized Bed Reactor in a Pilot Plant. Fuel Process. Technol. 2018, 179, 17–31. [Google Scholar] [CrossRef]
  110. Ly, H.V.; Kim, S.S.; Choi, J.H.; Woo, H.C.; Kim, J. Fast Pyrolysis of Saccharina Japonica Alga in a Fixed-Bed Reactor for Bio-Oil Production. Energy Convers. Manag. 2016, 122, 526–534. [Google Scholar] [CrossRef]
  111. Santos, J.; Ouadi, M.; Jahangiri, H.; Hornung, A. Integrated Intermediate Catalytic Pyrolysis of Wheat Husk. Food Bioprod. Process. 2019, 114, 23–30. [Google Scholar] [CrossRef]
  112. Jouhara, H.; Ahmad, D.; van den Boogaert, I.; Katsou, E.; Simons, S.; Spencer, N. Pyrolysis of Domestic Based Feedstock at Temperatures up to 300 °C. Therm. Sci. Eng. Prog. 2018, 5, 117–143. [Google Scholar] [CrossRef]
  113. Gong, X.; Kung, C.C.; Zhang, L. An Economic Evaluation on Welfare Distribution and Carbon Sequestration under Competitive Pyrolysis Technologies. Energy Explor. Exploit. 2021, 39, 553–570. [Google Scholar] [CrossRef]
  114. Lee, X.J.; Ong, H.C.; Gan, Y.Y.; Chen, W.H.; Mahlia, T.M.I. State of Art Review on Conventional and Advanced Pyrolysis of Macroalgae and Microalgae for Biochar, Bio-Oil and Bio-Syngas Production. Energy Convers. Manag. 2020, 210, 112707. [Google Scholar] [CrossRef]
  115. Singh, N.; Singhania, R.R.; Nigam, P.S.; di Dong, C.; Patel, A.K.; Puri, M. Global Status of Lignocellulosic Biorefinery: Challenges and Perspectives. Bioresour. Technol. 2022, 344, 126415. [Google Scholar] [CrossRef]
  116. Lee, X.J.; Lee, L.Y.; Gan, S.; Thangalazhy-Gopakumar, S.; Ng, H.K. Biochar Potential Evaluation of Palm Oil Wastes through Slow Pyrolysis: Thermochemical Characterization and Pyrolytic Kinetic Studies. Bioresour. Technol. 2017, 236, 155–163. [Google Scholar] [CrossRef]
  117. Brebu, M.; Vasile, C. Thermal Degradation of Lignin-A Review. Cellul. Chem. Technol. 2010, 44, 353–363. [Google Scholar]
  118. Quan, C.; Gao, N.; Song, Q. Pyrolysis of Biomass Components in a TGA and a Fixed-Bed Reactor: Thermochemical Behaviors, Kinetics, and Product Characterization. J. Anal. Appl. Pyrolysis 2016, 121, 84–92. [Google Scholar] [CrossRef]
  119. Luo, D.; Wang, L.; Nan, H.; Cao, Y.; Wang, H.; Kumar, T.V.; Wang, C. Phosphorus Adsorption by Functionalized Biochar: A Review. Environ. Chem. Lett. 2022, 21, 497–524. [Google Scholar] [CrossRef]
  120. Kang, Z.; Jia, X.; Zhang, Y.; Kang, X.; Ge, M.; Liu, D.; Wang, C.; He, Z. A Review on Application of Biochar in the Removal of Pharmaceutical Pollutants through Adsorption and Persulfate-Based AOPs. Sustainability 2022, 14, 10128. [Google Scholar] [CrossRef]
  121. Reza, M.S.; Azad, A.K.; Bakar, M.S.A.; Karim, M.R.; Sharifpur, M.; Taweekun, J. Evaluation of Thermochemical Characteristics and Pyrolysis of Fish Processing Waste for Renewable Energy Feedstock. Sustainability 2022, 14, 1203. [Google Scholar] [CrossRef]
  122. Xu, Y.; Hu, X.; Li, W.; Shi, Y. Preparation and Characterization of Bio-Oil from Biomass. In Progress in Biomass and Bioenergy Production; Shaukat, S., Ed.; InTech: Shanghai, China, 2011; pp. 197–222. ISBN 978-953-307-491-7. [Google Scholar]
  123. Alvarez, J.; Amutio, M.; Lopez, G.; Bilbao, J.; Olazar, M. Fast Co-Pyrolysis of Sewage Sludge and Lignocellulosic Biomass in a Conical Spouted Bed Reactor. Fuel 2015, 159, 810–818. [Google Scholar] [CrossRef]
  124. Saber, M.; Nakhshiniev, B.; Yoshikawa, K. A Review of Production and Upgrading of Algal Bio-Oil. Renew. Sustain. Energy Rev. 2016, 58, 918–930. [Google Scholar] [CrossRef]
  125. Lin, B.-J.; Chen, W.-H. Sugarcane Bagasse Pyrolysis in a Carbon Dioxide Atmosphere with Conventional and Microwave-Assisted Heating. Front. Energy Res. 2015, 3, 4. [Google Scholar] [CrossRef]
  126. Bridgwater, A.V. Review of Fast Pyrolysis of Biomass and Product Upgrading. Biomass Bioenergy 2012, 38, 68–94. [Google Scholar] [CrossRef]
  127. Demirbas, A. Biofuels Sources, Biofuel Policy, Biofuel Economy and Global Biofuel Projections. Energy Convers. Manag. 2008, 49, 2106–2116. [Google Scholar] [CrossRef]
  128. Motasemi, F.; Afzal, M.T. A Review on the Microwave-Assisted Pyrolysis Technique. Renew. Sustain. Energy Rev. 2013, 28, 317–330. [Google Scholar] [CrossRef]
  129. Yang, C.; Li, R.; Cui, C.; Liu, S.; Qiu, Q.; Ding, Y.; Wu, Y.; Zhang, B. Catalytic Hydroprocessing of Microalgae-Derived Biofuels: A Review. Green Chem. 2016, 18, 3684–3699. [Google Scholar] [CrossRef]
  130. Campanella, A.; Harold, M.P. Fast Pyrolysis of Microalgae in a Falling Solids Reactor: Effects of Process Variables and Zeolite Catalysts. Biomass Bioenergy 2012, 46, 218–232. [Google Scholar] [CrossRef]
  131. Aydin, Ö.; Nakajima, H.; Kitahara, T. Current and Temperature Distributions in-Situ Acquired by Electrode-Segmentation along a Microtubular Solid Oxide Fuel Cell Operating with Syngas. J. Power Sources 2015, 293, 1053–1061. [Google Scholar] [CrossRef]
  132. Samolada, M.C.; Stoicos, T.; Vasalos, I.A. An Investigation of the Factors Controlling the Pyrolysis Product Yield of Greek Wood Biomass in a Fluidized Bed. J. Anal. Appl. Pyrolysis 1990, 18, 127–141. [Google Scholar] [CrossRef]
  133. Rahman, A.A.; Sulaiman, F.; Abdullah, N. Effect of Temperature on Pyrolysis Product of Empty Fruit Bunches. In Proceedings of the AIP Conference Proceedings 1657, Tahara, Japan, 22–23 October 2015; p. 40011. [Google Scholar]
  134. Demirbas, A. Effect of Temperature on Pyrolysis Products from Biomass. Energy Sources Part A Recover. Util. Environ. Eff. 2007, 29, 329–336. [Google Scholar] [CrossRef]
  135. Uddin, M.N.; Techato, K.; Taweekun, J.; Rahman, M.M.; Rasul, M.G.; Mahlia, T.M.I.; Ashrafur, S.M. An Overview of Recent Developments in Biomass Pyrolysis Technologies. Energies 2018, 11, 3115. [Google Scholar] [CrossRef]
  136. Kan, T.; Strezov, V.; Evans, T.J. Lignocellulosic Biomass Pyrolysis: A Review of Product Properties and Effects of Pyrolysis Parameters. Renew. Sustain. Energy Rev. 2016, 57, 1126–1140. [Google Scholar] [CrossRef]
  137. Solís, S. Biochar Production by Biomass Pyrolysis for Various Applications. Master’s Thesis, Luleå University of Technology, Luleå, Sweden, 2016. [Google Scholar]
  138. Safdari, M.-S.; Amini, E.; Weise, D.R.; Fletcher, T.H. Heating Rate and Temperature Effects on Pyrolysis Products from Live Wildland Fuels. Fuel 2019, 242, 295–304. [Google Scholar] [CrossRef]
  139. Alsaleh, A.; Sattler, M.L. Waste Tire Pyrolysis: Influential Parameters and Product Properties. Curr. Sustain. Energy Rep. 2014, 1, 129–135. [Google Scholar] [CrossRef]
  140. Salehi, E.; Abedi, J.; Harding, T. Bio-Oil from Sawdust: Pyrolysis of Sawdust in a Fixed-Bed System. Energy Fuels 2009, 23, 3767–3772. [Google Scholar] [CrossRef]
  141. Ozbay, N.; Pütün, A.E.; Pütün, E. Bio-Oil Production from Rapid Pyrolysis of Cottonseed Cake: Product Yields and Compositions. Int. J. Energy Res. 2006, 30, 501–510. [Google Scholar] [CrossRef]
  142. Basu, P. Biomass Gasification, Pyrolysis and Torrefaction: Practical Design and Theory; Elsevier: Amsterdam, The Netherlands, 2018; ISBN 9780128129920. [Google Scholar]
  143. Mastral, F.J.; Esperanza, E.; García, P.; Juste, M.; Garciía, P.; Juste, M. Pyrolysis of High-Density Polyethylene in a Fluidised Bed Reactor. Influence of the Temperature and Residence Time. J. Anal. Appl. Pyrolysis 2002, 63, 1–15. [Google Scholar] [CrossRef]
  144. Boscagli, C.; Tomasi Morgano, M.; Raffelt, K.; Leibold, H.; Grunwaldt, J.D. Influence of Feedstock, Catalyst, Pyrolysis and Hydrotreatment Temperature on the Composition of Upgraded Oils from Intermediate Pyrolysis. Biomass Bioenergy 2018, 116, 236–248. [Google Scholar] [CrossRef]
  145. Azeta, O.; Ayeni, A.O.; Agboola, O.; Elehinafe, F.B. A Review on the Sustainable Energy Generation from the Pyrolysis of Coconut Biomass. Sci. African 2021, 13, e00909. [Google Scholar] [CrossRef]
  146. Solar, J.; de Marco, I.; Caballero, B.M.; Lopez-Urionabarrenechea, A.; Rodriguez, N.; Agirre, I.; Adrados, A. Influence of Temperature and Residence Time in the Pyrolysis of Woody Biomass Waste in a Continuous Screw Reactor. Biomass Bioenergy 2016, 95, 416–423. [Google Scholar] [CrossRef]
  147. Luo, S.; Xiao, B.; Hu, Z.; Liu, S. Effect of Particle Size on Pyrolysis of Single-Component Municipal Solid Waste in Fixed Bed Reactor. Int. J. Hydrogen Energy 2010, 35, 93–97. [Google Scholar] [CrossRef]
  148. Martínez, J.D.; Puy, N.; Murillo, R.; García, T.; Navarro, M.V.; Mastral, A.M. Waste Tyre Pyrolysis—A Review. Renew. Sustain. Energy Rev. 2013, 23, 179–213. [Google Scholar] [CrossRef]
  149. Aguilar, G.; Muley, P.D.; Henkel, C.; Boldor, D. Effects of Biomass Particle Size on Yield and Composition of Pyrolysis Bio-Oil Derived from Chinese Tallow Tree (Triadica sebifera L.) and Energy Cane (Saccharum complex) in an Inductively Heated Reactor. AIMS Energy 2015, 3, 838–850. [Google Scholar] [CrossRef]
  150. Bridgwater, A. Fast Pyrolysis of Biomass for the Production of Liquids. In Biomass Combustion Science, Technology and Engineering; Woodhead Publishing: Sawston, UK, 2013; pp. 130–171. ISBN 9780857091314. [Google Scholar]
  151. Ibrahim, H.A.-H. Introductory Chapter: Pyrolysis. In Recent Advances in Pyrolysis; IntechOpen: London, UK, 2020; ISBN 978-1-78984-064-3. [Google Scholar]
  152. Qiu, B.; Tao, X.; Wang, J.; Liu, Y.; Li, S.; Chu, H. Research Progress in the Preparation of High-Quality Liquid Fuels and Chemicals by Catalytic Pyrolysis of Biomass: A Review. Energy Convers. Manag. 2022, 261, 115647. [Google Scholar] [CrossRef]
  153. Kakaei, K.; Esrafili, M.D.; Ehsani, A. Introduction to Catalysis. Interface Sci. Technol. 2019, 27, 1–21. [Google Scholar] [CrossRef]
  154. Fermoso, J.; Coronado, J.M.; Serrano, D.P.; Pizarro, P. Pyrolysis of Microalgae for Fuel Production. In Microalgae-Based Biofuels and Bioproducts: From Feedstock Cultivation to End-Products; Elsevier Inc.: Duxford, UK, 2017; pp. 259–281. ISBN 9780081010273. [Google Scholar]
  155. Panda, A.K.; Singh, R.K.; Mishra, D.K. Thermolysis of Waste Plastics to Liquid fuelA Suitable Method for Plastic Waste Management and Manufacture of Value Added products—A World Prospective. Renew. Sustain. Energy Rev. 2010, 14, 233–248. [Google Scholar] [CrossRef]
  156. Yadav, K.; Jagadevan, S. Influence of Process Parameters on Synthesis of Biochar by Pyrolysis of Biomass: An Alternative Source of Energy. In Recent Advances in Pyrolysis; IntechOpen: London, UK, 2020; pp. 1–14. [Google Scholar] [CrossRef]
  157. Torri, C.; Reinikainen, M.; Lindfors, C.; Fabbri, D.; Oasmaa, A.; Kuoppala, E. Investigation on Catalytic Pyrolysis of Pine Sawdust: Catalyst Screening by Py-GC-MIP-AED. J. Anal. Appl. Pyrolysis 2010, 88, 7–13. [Google Scholar] [CrossRef]
  158. Abu Bakar, M.S.; Titiloye, J.O. Catalytic Pyrolysis of Rice Husk for Bio-Oil Production. J. Anal. Appl. Pyrolysis 2013, 103, 362–368. [Google Scholar] [CrossRef]
  159. French, R.; Czernik, S. Catalytic Pyrolysis of Biomass for Biofuels Production. Fuel Process. Technol. 2010, 91, 25–32. [Google Scholar] [CrossRef]
  160. Thangalazhy-Gopakumar, S.; Adhikari, S.; Gupta, R.B.; Tu, M.; Taylor, S. Production of Hydrocarbon Fuels from Biomass Using Catalytic Pyrolysis under Helium and Hydrogen Environments. Bioresour. Technol. 2011, 102, 6742–6749. [Google Scholar] [CrossRef]
  161. Güngör, A.; Önenç, S.; Uçar, S.; Yanik, J. Comparison between the “one-Step” and “two-Step” Catalytic Pyrolysis of Pine Bark. J. Anal. Appl. Pyrolysis 2012, 97, 39–48. [Google Scholar] [CrossRef]
  162. Li, J.; Li, X.; Zhou, G.; Wang, W.; Wang, C.; Komarneni, S.; Wang, Y. Catalytic Fast Pyrolysis of Biomass with Mesoporous ZSM-5 Zeolites Prepared by Desilication with NaOH Solutions. Appl. Catal. A Gen. 2014, 470, 115–122. [Google Scholar] [CrossRef]
  163. Stefanidis, S.D.D.; Kalogiannis, K.G.G.; Iliopoulou, E.F.F.; Lappas, A.A.A.; Pilavachi, P.A.A. In-Situ Upgrading of Biomass Pyrolysis Vapors: Catalyst Screening on a Fixed Bed Reactor. Bioresour. Technol. 2011, 102, 8261–8267. [Google Scholar] [CrossRef]
  164. Pan, P.; Hu, C.; Yang, W.; Li, Y.; Dong, L.; Zhu, L.; Tong, D.; Qing, R.; Fan, Y. The Direct Pyrolysis and Catalytic Pyrolysis of Nannochloropsis Sp. Residue for Renewable Bio-Oils. Bioresour. Technol. 2010, 101, 4593–4599. [Google Scholar] [CrossRef]
  165. Zhang, H.; Xiao, R.; Huang, H.; Xiao, G. Comparison of Non-Catalytic and Catalytic Fast Pyrolysis of Corncob in a Fluidized Bed Reactor. Bioresour. Technol. 2009, 100, 1428–1434. [Google Scholar] [CrossRef]
  166. Compton, D.L.; Jackson, M.A.; Mihalcik, D.J.; Mullen, C.A.; Boateng, A.A. Catalytic Pyrolysis of Oak via Pyroprobe and Bench Scale, Packed Bed Pyrolysis Reactors. J. Anal. Appl. Pyrolysis 2011, 90, 174–181. [Google Scholar] [CrossRef]
  167. Demiral, İ.; Şensöz, S. The Effects of Different Catalysts on the Pyrolysis of Industrial Wastes (Olive and Hazelnut Bagasse). Bioresour. Technol. 2008, 99, 8002–8007. [Google Scholar] [CrossRef]
  168. Park, H.J.; Heo, H.S.; Park, Y.-K.; Yim, J.-H.; Jeon, J.-K.; Park, J.; Ryu, C.; Kim, S.-S. Clean Bio-Oil Production from Fast Pyrolysis of Sewage Sludge: Effects of Reaction Conditions and Metal Oxide Catalysts. Bioresour. Technol. 2010, 101, S83–S85. [Google Scholar] [CrossRef]
  169. Zhou, L.; Yang, H.; Wu, H.; Wang, M.; Cheng, D. Catalytic Pyrolysis of Rice Husk by Mixing with Zinc Oxide: Characterization of Bio-Oil and Its Rheological Behavior. Fuel Process. Technol. 2013, 106, 385–391. [Google Scholar] [CrossRef]
  170. Babich, I.V.; van der Hulst, M.; Lefferts, L.; Moulijn, J.A.; O’Connor, P.; Seshan, K. Catalytic Pyrolysis of Microalgae to High-Quality Liquid Bio-Fuels. Biomass Bioenergy 2011, 35, 3199–3207. [Google Scholar] [CrossRef]
  171. Rutkowski, P. Pyrolysis of Cellulose, Xylan and Lignin with the K2CO3 and ZnCl2 Addition for Bio-Oil Production. Fuel Process. Technol. 2011, 92, 517–522. [Google Scholar] [CrossRef]
  172. Barbooti, M.M.; Matlub, F.K.; Hadi, H.M. Catalytic Pyrolysis of Phragmites (Reed): Investigation of Its Potential as a Biomass Feedstock. J. Anal. Appl. Pyrolysis 2012, 98, 1–6. [Google Scholar] [CrossRef]
  173. Wolff, A. Management of Waste from the Health-Care Sector, 1st ed.; Elsevier B.V.: Amsterdam, The Netherlands, 2018; Volume 11, ISBN 9780444638571. [Google Scholar]
  174. Jerzak, W.; Gao, N.; Kalemba-Rec, I.; Magdziarz, A. Catalytic Intermediate Pyrolysis of Post-Extraction Rapeseed Meal by Reusing ZSM-5 and Zeolite Y Catalysts. Catal. Today 2022, 404, 63–77. [Google Scholar] [CrossRef]
  175. Zadeh, Z.E.; Abdulkhani, A.; Aboelazayem, O.; Saha, B. Recent Insights into Lignocellulosic Biomass Pyrolysis: A Critical Review on Pretreatment, Characterization, and Products Upgrading. Processes 2020, 8, 799. [Google Scholar] [CrossRef]
  176. Luo, G.; Resende, F.L.P. In-Situ and Ex-Situ Upgrading of Pyrolysis Vapors from Beetle-Killed Trees. Fuel 2016, 166, 367–375. [Google Scholar] [CrossRef]
  177. Muneer, B.; Zeeshan, M.; Qaisar, S.; Razzaq, M.; Iftikhar, H. Influence of in-Situ and Ex-Situ HZSM-5 Catalyst on Co-Pyrolysis of Corn Stalk and Polystyrene with a Focus on Liquid Yield and Quality. J. Clean. Prod. 2019, 237, 117762. [Google Scholar] [CrossRef]
  178. Kabakcı, S.B.; Hacıbektaşoğlu, Ş. Catalytic Pyrolysis of Biomass Literature Review. In Pyrolysis; Samer, M., Ed.; IntechOpen Limited: London, UK, 2017; pp. 167–196. [Google Scholar]
  179. Wan, S.; Wang, Y. A Review on Ex Situ Catalytic Fast Pyrolysis of Biomass. Front. Chem. Sci. Eng. 2014, 8, 280–294. [Google Scholar] [CrossRef]
  180. Iliopoulou, E.F.; Triantafyllidis, K.S.; Lappas, A.A. Overview of Catalytic Upgrading of Biomass Pyrolysis Vapors toward the Production of Fuels and High-Value Chemicals. Wiley Interdiscip. Rev. Energy Environ. 2019, 8, e322. [Google Scholar] [CrossRef]
  181. Wang, C.; Lei, H.; Zou, R.; Qian, M.; Mateo, W.; Lin, X.; Ruan, R. Biochar-Driven Simplification of the Compositions of Cellulose-Pyrolysis-Derived Biocrude Oil Coupled with the Promotion of Hydrogen Generation. Bioresour. Technol. 2021, 334, 125251. [Google Scholar] [CrossRef]
  182. Abu Bakar, M. Catalytic Intermediate Pyrolysis of Brunei Rice Husk for Bio-Oil Production. Ph.D. Thesis, Aston University, Birmingham, UK, 2013. [Google Scholar]
  183. Yang, H.; Coolman, R.; Karanjkar, P.; Wang, H.; Dornath, P.; Chen, H.; Fan, W.; Conner, W.C.; Mountziaris, T.J.; Huber, G. The Effects of Contact Time and Coking on the Catalytic Fast Pyrolysis of Cellulose. Green Chem. 2017, 19, 286–297. [Google Scholar] [CrossRef]
  184. Li, X.; Su, L.; Wang, Y.; Yu, Y.; Wang, C.; Li, X.; Wang, Z. Catalytic Fast Pyrolysis of Kraft Lignin with HZSM-5 Zeolite for Producing Aromatic Hydrocarbons. Front. Environ. Sci. Eng. China 2012, 6, 295–303. [Google Scholar] [CrossRef]
  185. Argyle, M.D.; Bartholomew, C.H. Heterogeneous Catalyst Deactivation and Regeneration: A Review. Catalysts 2015, 5, 145–269. [Google Scholar] [CrossRef]
  186. Carlson, T.R.; Tompsett, G.A.; Conner, W.C.; Huber, G.W. Aromatic Production from Catalytic Fast Pyrolysis of Biomass-Derived Feedstocks. Top. Catal. 2009, 52, 241–252. [Google Scholar] [CrossRef]
  187. Mebrahtu, C.; Krebs, F.; Abate, S.; Perathoner, S.; Centi, G.; Palkovits, R. CO2 Methanation: Principles and Challenges. In Studies in Surface Science and Catalysis; Elsevier Inc.: Amsterdam, The Netherlands, 2019; Volume 178, pp. 85–103. [Google Scholar]
  188. Hervy, M.; Weiss-Hortala, E.; Pham Minh, D.; Dib, H.; Villot, A.; Gérente, C.; Berhanu, S.; Chesnaud, A.; Thorel, A.; Le Coq, L.; et al. Reactivity and Deactivation Mechanisms of Pyrolysis Chars from Bio-Waste during Catalytic Cracking of Tar. Appl. Energy 2019, 237, 487–499. [Google Scholar] [CrossRef]
  189. Zhang, X.; Lei, H.; Zhu, L.; Qian, M.; Chan, J.C.; Zhu, X.; Liu, Y.; Yadavalli, G.; Yan, D.; Wang, L.; et al. Development of a Catalytically Green Route from Diverse Lignocellulosic Biomasses to High-Density Cycloalkanes for Jet Fuels. Catal. Sci. Technol. 2016, 6, 4210–4220. [Google Scholar] [CrossRef]
  190. Onwudili, J.A.; Muhammad, C.; Williams, P.T. Influence of Catalyst Bed Temperature and Properties of Zeolite Catalysts on Pyrolysis-Catalysis of a Simulated Mixed Plastics Sample for the Production of Upgraded Fuels and Chemicals. J. Energy Inst. 2019, 92, 1337–1347. [Google Scholar] [CrossRef]
  191. Zhou, G.; Jensen, P.A.; Le, D.M.; Knudsen, N.O.; Jensen, A.D. Direct Upgrading of Fast Pyrolysis Lignin Vapor over the HZSM-5 Catalyst. Green Chem. 2016, 18, 1965–1975. [Google Scholar] [CrossRef]
  192. Williams, P.T.; Nugranad, N. Comparison of Products from the Pyrolysis and Catalytic Pyrolysis of Rice Husks. Energy 2000, 25, 493–513. [Google Scholar] [CrossRef]
  193. Al-Salem, S.M.; Antelava, A.; Constantinou, A.; Manos, G.; Dutta, A. A Review on Thermal and Catalytic Pyrolysis of Plastic Solid Waste (PSW). J. Environ. Manag. 2017, 197, 177–198. [Google Scholar] [CrossRef] [PubMed]
  194. Rahman, M.M.; Liu, R.; Cai, J. Catalytic Fast Pyrolysis of Biomass over Zeolites for High Quality Bio-Oil—A Review. Fuel Process. Technol. 2018, 180, 32–46. [Google Scholar] [CrossRef]
  195. Martin, A. Zeolite Catalysis. Appl. Catal. 1990, 59, 118. [Google Scholar] [CrossRef]
  196. Yu, Y.; Li, X.; Su, L.; Zhang, Y.; Wang, Y.; Zhang, H. The Role of Shape Selectivity in Catalytic Fast Pyrolysis of Lignin with Zeolite Catalysts. Appl. Catal. A Gen. 2012, 447–448, 115–123. [Google Scholar] [CrossRef]
  197. Gao, L.; Sun, J.; Xu, W.; Xiao, G. Catalytic Pyrolysis of Natural Algae over Mg-Al Layered Double oxides/ZSM-5 (MgAl-LDO/ZSM-5) for Producing Bio-Oil with Low Nitrogen Content. Bioresour. Technol. 2017, 225, 293–298. [Google Scholar] [CrossRef]
  198. Galadima, A.; Muraza, O. In Situ Fast Pyrolysis of Biomass with Zeolite Catalysts for Bioaromatics/gasoline Production: A Review. Energy Convers. Manag. 2015, 105, 338–354. [Google Scholar] [CrossRef]
  199. Shirazi, L.; Jamshidi, E.; Ghasemi, M.R. The Effect of Si/Al Ratio of ZSM-5 Zeolite on Its Morphology, Acidity and Crystal Size. Cryst. Res. Technol. 2008, 43, 1300–1306. [Google Scholar] [CrossRef]
  200. Speight, J.G. Upgrading by Hydrocracking; Elsevier: Amsterdam, The Netherlands, 2019; pp. 467–528. [Google Scholar]
  201. Rehan, M.; Miandad, R.; Barakat, M.A.; Ismail, I.M.I.; Almeelbi, T.; Gardy, J.; Hassanpour, A.; Khan, M.Z.; Demirbas, A.; Nizami, A.S. Effect of Zeolite Catalysts on Pyrolysis Liquid Oil. Int. Biodeterior. Biodegrad. 2017, 119, 162–175. [Google Scholar] [CrossRef]
  202. Veses, A.; Aznar, M.; López, J.M.; Callén, M.S.; Murillo, R.; García, T. Production of Upgraded Bio-Oils by Biomass Catalytic Pyrolysis in an Auger Reactor Using Low Cost Materials. Fuel 2015, 141, 17–22. [Google Scholar] [CrossRef]
  203. Ly, H.V.; Lim, D.-H.; Sim, J.W.; Kim, S.-S.; Kim, J. Catalytic Pyrolysis of Tulip Tree (Liriodendron) in Bubbling Fluidized-Bed Reactor for Upgrading Bio-Oil Using Dolomite Catalyst. Energy 2018, 162, 564–575. [Google Scholar] [CrossRef]
  204. Mehmood, M. Dolomite and Dolomitization Model—A Short Review. Int. J. Hydrol. 2018, 2, 1–5. [Google Scholar] [CrossRef]
  205. Maisano, S.; Urbani, F.; Mondello, N.; Chiodo, V. Catalytic Pyrolysis of Mediterranean Sea Plant for Bio-Oil Production. Int. J. Hydrog. Energy 2017, 42, 28082–28092. [Google Scholar] [CrossRef]
  206. Valle, B.; Aramburu, B.; Santiviago, C.; Bilbao, J.; Gayubo, A.G. Upgrading of Bio-Oil in a Continuous Process with Dolomite Catalyst. Energy Fuels 2014, 28, 6419–6428. [Google Scholar] [CrossRef]
  207. Tursunov, O. A Comparison of Catalysts Zeolite and Calcined Dolomite for Gas Production from Pyrolysis of Municipal Solid Waste (MSW). Ecol. Eng. 2014, 69, 237–243. [Google Scholar] [CrossRef]
  208. Gerçel, H.F. Bio-Oil Production from Onopordum Acanthium L. by Slow Pyrolysis. J. Anal. Appl. Pyrolysis 2011, 92, 233–238. [Google Scholar] [CrossRef]
  209. Sulman, M.; Kosivtsov, Y.; Sulman, E.; Alfyorov, V.; Lugovoy, Y.; Molchanov, V.; Tyamina, I.; Misnikov, O.; Afanasjev, A.; Kumar, N.; et al. Influence of Aluminosilicate Materials on the Peat Low-Temperature Pyrolysis and Gas Formation. Chem. Eng. J. 2009, 154, 355–360. [Google Scholar] [CrossRef]
  210. Suprianto, T.; Winarto; Wijayanti, W.; Wardana, I. Effect of Activated Carbon Catalyst on the Cracking of Biomass Molecules into Light Hydrocarbons in Biomass Pyrolysis. In IOP Conference Series: Materials Science and Engineering, Proceedings of the 2nd International Conference on Mechanical Engineering Research and Application (iCOMERA 2020), Malang, Indonesia, 7–9 October 2020; IOPScience: Bristol, UK, 2021; Volume 1034, p. 12079. [Google Scholar]
  211. Zhang, Y.; Lei, H.; Yang, Z.; Duan, D.; Villota, E.; Ruan, R. From Glucose-Based Carbohydrates to Phenol-Rich Bio-Oils Integrated with Syngas Production via Catalytic Pyrolysis over an Activated Carbon Catalyst. Green Chem. 2018, 20, 3346–3358. [Google Scholar] [CrossRef]
  212. Duan, D.; Zhang, Y.; Lei, H.; Villota, E.; Ruan, R. Renewable Jet-Fuel Range Hydrocarbons Production from Co-Pyrolysis of Lignin and Soapstock with the Activated Carbon Catalyst. Waste Manag. 2019, 88, 1–9. [Google Scholar] [CrossRef]
  213. Bu, Q.; Lei, H.; Wang, L.; Wei, Y.; Zhu, L.; Liu, Y.; Liang, J.; Tang, J. Renewable Phenols Production by Catalytic Microwave Pyrolysis of Douglas Fir Sawdust Pellets with Activated Carbon Catalysts. Bioresour. Technol. 2013, 142, 546–552. [Google Scholar] [CrossRef] [PubMed]
  214. Bu, Q.; Lei, H.; Wang, L.; Yadavalli, G.; Wei, Y.; Zhang, X.; Zhu, L.; Liu, Y. Biofuel Production from Catalytic Microwave Pyrolysis of Douglas Fir Pellets over Ferrum-Modified Activated Carbon Catalyst. J. Anal. Appl. Pyrolysis 2015, 112, 74–79. [Google Scholar] [CrossRef]
  215. Oxide—Wikipedia. Available online: https://en.wikipedia.org/wiki/Oxide (accessed on 4 May 2021).
  216. Shi, K.; Shao, S.; Huang, Q.; Liang, X.; Lan, J.; Ya, L. Review of Catalytic Pyrolysis of Biomass for Bio-Oil. In Proceedings of the ICMREE2011—Proceedings 2011 International Conference on Materials for Renewable Energy and Environment, Shanghai, China, 20–22 May 2011; Volume 1, pp. 317–321. [Google Scholar]
  217. Liu, C.; Wang, H.; Karim, A.M.; Sun, J.; Wang, Y. Catalytic Fast Pyrolysis of Lignocellulosic Biomass. Chem. Soc. Rev. 2014, 43, 7594–7623. [Google Scholar] [CrossRef]
  218. Oxide|Chemical Compound|Britannica. Available online: https://www.britannica.com/science/oxide (accessed on 12 April 2023).
  219. Zhang, C.; Hu, X.; Guo, H.; Wei, T.; Dong, D.; Hu, G.; Hu, S.; Xiang, J.; Liu, Q.; Wang, Y. Pyrolysis of Poplar, Cellulose and Lignin: Effects of Acidity and Alkalinity of the Metal Oxide Catalysts. J. Anal. Appl. Pyrolysis 2018, 134, 590–605. [Google Scholar] [CrossRef]
  220. Che, Q.; Yang, M.; Wang, X.; Chen, X.; Chen, W.; Yang, Q.; Yang, H.; Chen, H. Aromatics Production with Metal Oxides and ZSM-5 as Catalysts in Catalytic Pyrolysis of Wood Sawdust. Fuel Process. Technol. 2019, 188, 146–152. [Google Scholar] [CrossRef]
  221. Li, D.; Yao, F.; Guo, Q.X. Upgraded Acidic Components of Bio-Oil through Catalytic Ketonic Condensation. Energy Fuels 2009, 23, 564–568. [Google Scholar] [CrossRef]
  222. Aysu, T. Pyrolysis of Isochrysis Microalgae with Metal Oxide Catalysts for Bio-Oil Production. J. Turk. Chem. Soc. Sect. A Chem. 2016, 4, 395. [Google Scholar] [CrossRef]
  223. Chong, Y.Y.; Thangalazhy-Gopakumar, S.; Ng, H.K.; Lee, L.Y.; Gan, S. Effect of Oxide Catalysts on the Properties of Bio-Oil from in-Situ Catalytic Pyrolysis of Palm Empty Fruit Bunch Fiber. J. Environ. Manag. 2019, 247, 38–45. [Google Scholar] [CrossRef] [PubMed]
  224. Zheng, Y.; Wang, J.; Wang, D.; Zheng, Z. Advanced Catalytic Upgrading of Biomass Pyrolysis Vapor to Bio-Aromatics Hydrocarbon: A Review. Appl. Energy Combust. Sci. 2022, 10, 100061. [Google Scholar] [CrossRef]
  225. Zheng, Y.; Wang, J.; Liu, C.; Lu, Y.; Lin, X.; Li, W.; Zheng, Z. Efficient and Stable Ni-Cu Catalysts for Ex Situ Catalytic Pyrolysis Vapor Upgrading of Oleic Acid into Hydrocarbon: Effect of Catalyst Support, Process Parameters and Ni-to-Cu Mixed Ratio. Renew. Energy 2020, 154, 797–812. [Google Scholar] [CrossRef]
  226. Jiang, L.; Wang, Y.; Dai, L.; Yu, Z.; Yang, Q.; Yang, S.; Jiang, D.; Ma, Z.; Wu, Q.; Zhang, B.; et al. Co-Pyrolysis of Biomass and Soapstock in a Downdraft Reactor Using a Novel ZSM-5/SiC Composite Catalyst. Bioresour. Technol. 2019, 279, 202–208. [Google Scholar] [CrossRef]
  227. Kaewpengkrow, P.; Atong, D.; Sricharoenchaikul, V. Selective Catalytic Fast Pyrolysis of Jatropha Curcas Residue with Metal Oxide Impregnated Activated Carbon for Upgrading Bio-Oil. Int. J. Hydrogen Energy 2017, 42, 18397–18409. [Google Scholar] [CrossRef]
  228. Nikkhah, H.; Tavasoli, A.; Jafarian, S. Investigating the Influence of Acid Washing Pretreatment and Zn/activated Biochar Catalyst on Thermal Conversion of Cladophora Glomerata to Value-Added Bio-Products. Energy Convers. Manag. 2020, 225, 113392. [Google Scholar] [CrossRef]
  229. Mohamed, B.A.; Kim, C.S.; Ellis, N.; Bi, X. Microwave-Assisted Catalytic Pyrolysis of Switchgrass for Improving Bio-Oil and Biochar Properties. Bioresour. Technol. 2016, 201, 121–132. [Google Scholar] [CrossRef] [PubMed]
  230. Vispute, T. Pyrolysis Oils: Characterization, Stability Analysis, and Catalytic Upgrading to Fuels and Chemicals. Ph.D. Thesis, University of Massachusetts Amherst, Amherst, MA, USA, 2011. [Google Scholar]
  231. Thangalazhy-Gopakumar, S.; Adhikari, S.; Ravindran, H.; Gupta, R.B.; Fasina, O.; Tu, M.; Fernando, S.D. Physiochemical Properties of Bio-Oil Produced at Various Temperatures from Pine Wood Using an Auger Reactor. Bioresour. Technol. 2010, 101, 8389–8395. [Google Scholar] [CrossRef]
  232. Chen, X.; Chen, Y.; Yang, H.; Wang, X.; Che, Q.; Chen, W.; Chen, H. Catalytic Fast Pyrolysis of Biomass: Selective Deoxygenation to Balance the Quality and Yield of Bio-Oil. Bioresour. Technol. 2019, 273, 153–158. [Google Scholar] [CrossRef]
  233. Veses, A.; Aznar, M.; Martínez, I.; Martínez, J.D.; López, J.M.; Navarro, M.V.; Callén, M.S.; Murillo, R.; García, T. Catalytic Pyrolysis of Wood Biomass in an Auger Reactor Using Calcium-Based Catalysts. Bioresour. Technol. 2014, 162, 250–258. [Google Scholar] [CrossRef]
  234. Thangalazhy-Gopakumar, S.; Wei Lee, C.; Gan, S.; Kiat Ng, H.; Yee Lee, L. Comparison of Bio-Oil Properties from Non-Catalytic and In-Situ Catalytic Fast Pyrolysis of Palm Empty Fruit Bunch. Mater. Today Proc. 2018, 5, 23456–23465. [Google Scholar] [CrossRef]
  235. Muley, P.D.; Henkel, C.; Abdollahi, K.K.; Boldor, D. Pyrolysis and Catalytic Upgrading of Pinewood Sawdust Using an Induction Heating Reactor. Energy Fuels 2015, 29, 7375–7385. [Google Scholar] [CrossRef]
  236. Sheng, C.; Azevedo, J.L.T. Estimating the Higher Heating Value of Biomass Fuels from Basic Analysis Data. Biomass Bioenergy 2005, 28, 499–507. [Google Scholar] [CrossRef]
  237. Pulidori, E.; Gonzalez-Rivera, J.; Pelosi, C.; Ferrari, C.; Bernazzani, L.; Bramanti, E.; Tiné, M.R.; Duce, C.; Pulidori, E.; Gonzalez-Rivera, J.; et al. Thermochemical Evaluation of Different Waste Biomasses (Citrus Peels, Aromatic Herbs, and Poultry Feathers) towards Their Use for Energy Production. Thermo 2023, 3, 66–75. [Google Scholar] [CrossRef]
  238. Ben, H.; Wu, F.; Wu, Z.; Han, G.; Jiang, W.; Ragauskas, A.J. A Comprehensive Characterization of Pyrolysis Oil from Softwood Barks. Polymers 2019, 11, 1387. [Google Scholar] [CrossRef] [PubMed]
  239. Aysu, T.; Durak, H.; Güner, S.; Bengü, A.Ş.; Esim, N. Bio-Oil Production via Catalytic Pyrolysis of Anchusa Azurea: Effects of Operating Conditions on Product Yields and Chromatographic Characterization. Bioresour. Technol. 2016, 205, 7–14. [Google Scholar] [CrossRef]
  240. Zabeti, M.; Nguyen, T.S.; Lefferts, L.; Heeres, H.J.; Seshan, K. In Situ Catalytic Pyrolysis of Lignocellulose Using Alkali-Modified Amorphous Silica Alumina. Bioresour. Technol. 2012, 118, 374–381. [Google Scholar] [CrossRef] [PubMed]
  241. Yu, F.; Gao, L.; Wang, W.; Zhang, G.; Ji, J. Bio-Fuel Production from the Catalytic Pyrolysis of Soybean Oil over Me-Al-MCM-41 (Me = La, Ni or Fe) Mesoporous Materials. J. Anal. Appl. Pyrolysis 2013, 104, 325–329. [Google Scholar] [CrossRef]
  242. Nokkosmäki, M.I.; Kuoppala, E.T.; Leppämäki, E.A.; Krause, A.O.I. Catalytic Conversion of Biomass Pyrolysis Vapours with Zinc Oxide. J. Anal. Appl. Pyrolysis 2000, 55, 119–131. [Google Scholar] [CrossRef]
  243. Mante, O.D.; Agblevor, F.A. Catalytic Conversion of Biomass to Bio-Syncrude Oil. Biomass Convers. Biorefinery 2011, 1, 203–215. [Google Scholar] [CrossRef]
  244. Dong, Z.; Bai, X.; Xu, D.; Li, W. Machine Learning Prediction of Pyrolytic Products of Lignocellulosic Biomass Based on Physicochemical Characteristics and Pyrolysis Conditions. Bioresour. Technol. 2023, 367, 128182. [Google Scholar] [CrossRef]
  245. Bal, M.; Amasyali, M.F.; Sever, H.; Kose, G.; Demirhan, A. Performance Evaluation of the Machine Learning Algorithms Used in Inference Mechanism of a Medical Decision Support System. Sci. World J. 2014, 2014, 137896. [Google Scholar] [CrossRef]
  246. Wang, Z.; Peng, X.; Xia, A.; Shah, A.A.; Huang, Y.; Zhu, X.; Zhu, X.; Liao, Q. The Role of Machine Learning to Boost the Bioenergy and Biofuels Conversion. Bioresour. Technol. 2022, 343, 126099. [Google Scholar] [CrossRef] [PubMed]
  247. Ascher, S.; Watson, I.; You, S. Machine Learning Methods for Modelling the Gasification and Pyrolysis of Biomass and Waste. Renew. Sustain. Energy Rev. 2022, 155, 111902. [Google Scholar] [CrossRef]
  248. Pan, R.; Martins, M.F.; Debenest, G. Optimization of Oil Production through Ex-Situ Catalytic Pyrolysis of Waste Polyethylene with Activated Carbon. Energy 2022, 248, 123514. [Google Scholar] [CrossRef]
  249. Yang, K.; Wu, K.; Zhang, H. Machine Learning Prediction of the Yield and Oxygen Content of Bio-Oil via Biomass Characteristics and Pyrolysis Conditions. Energy 2022, 254, 124320. [Google Scholar] [CrossRef]
  250. Balsora, H.K.; Kartik, A.; Dua, V.; Joshi, J.B.; Kataria, G.; Sharma, A.; Chakinala, A.G. Machine Learning Approach for the Prediction of Biomass Pyrolysis Kinetics from Preliminary Analysis. J. Environ. Chem. Eng. 2022, 10, 108025. [Google Scholar] [CrossRef]
  251. Sharma, A.; Suryawanshi, B.; Mohanty, B.; Sawarkar, A.N. Comparison of Artificial Neural Network and Response Surface Methodology for Evaluation of the Predictive Capability of Bio-Oil Yield from Pyrolysis of Mangifera Indica Wood Sawdust. Fuel 2023, 338, 127251. [Google Scholar] [CrossRef]
  252. Anex, R.P.; Aden, A.; Kazi, F.K.; Fortman, J.; Swanson, R.M.; Wright, M.M.; Satrio, J.A.; Brown, R.C.; Daugaard, D.E.; Platon, A.; et al. Techno-Economic Comparison of Biomass-to-Transportation Fuels via Pyrolysis, Gasification, and Biochemical Pathways. Fuel 2010, 89, S29–S35. [Google Scholar] [CrossRef]
  253. Kumar, A.; Saini, K.; Bhaskar, T. Hydochar and Biochar: Production, Physicochemical Properties and Techno-Economic Analysis. Bioresour. Technol. 2020, 310, 123442. [Google Scholar] [CrossRef] [PubMed]
  254. López-González, D.; Puig-Gamero, M.; Acién, F.G.; García-Cuadra, F.; Valverde, J.L.; Sanchez-Silva, L. Energetic, Economic and Environmental Assessment of the Pyrolysis and Combustion of Microalgae and Their Oils. Renew. Sustain. Energy Rev. 2015, 51, 1752–1770. [Google Scholar] [CrossRef]
  255. Hoang, A.T.; Ong, H.C.; Fattah, I.M.R.; Chong, C.T.; Cheng, C.K.; Sakthivel, R.; Ok, Y.S. Progress on the Lignocellulosic Biomass Pyrolysis for Biofuel Production toward Environmental Sustainability. Fuel Process. Technol. 2021, 223, 106997. [Google Scholar] [CrossRef]
  256. Haeldermans, T.; Campion, L.; Kuppens, T.; Vanreppelen, K.; Cuypers, A.; Schreurs, S. A Comparative Techno-Economic Assessment of Biochar Production from Different Residue Streams Using Conventional and Microwave Pyrolysis. Bioresour. Technol. 2020, 318, 124083. [Google Scholar] [CrossRef]
  257. Zein, S.H.; Antony, A. Techno-Economic Analysis and Feasibility of Industrial-Scale Activated Carbon Production from Agricultural Pea Waste Using Microwave-Assisted Pyrolysis: A Circular Economy Approach. Processes 2022, 10, 1702. [Google Scholar] [CrossRef]
  258. Makepa, D.C.; Chihobo, C.H.; Ruziwa, W.R.; Musademba, D. A Systematic Review of the Techno-Economic Assessment and Biomass Supply Chain Uncertainties of Biofuels Production from Fast Pyrolysis of Lignocellulosic Biomass. Fuel Commun. 2023, 14, 100086. [Google Scholar] [CrossRef]
  259. Nematian, M.; Keske, C.; Ng’ombe, J.N. A Techno-Economic Analysis of Biochar Production and the Bioeconomy for Orchard Biomass. Waste Manag. 2021, 135, 467–477. [Google Scholar] [CrossRef] [PubMed]
  260. Kung, C.C.; Zhang, N. Renewable Energy from Pyrolysis Using Crops and Agricultural Residuals: An Economic and Environmental Evaluation. Energy 2015, 90, 1532–1544. [Google Scholar] [CrossRef]
  261. Thilakaratne, R.; Brown, T.; Li, Y.; Hu, G.; Brown, R. Mild Catalytic Pyrolysis of Biomass for Production of Transportation Fuels: A Techno-Economic Analysis. Green Chem. 2014, 16, 627–636. [Google Scholar] [CrossRef]
  262. Shoaib Ahmed Khan, M.; Grioui, N.; Halouani, K.; Benelmir, R. Techno-Economic Analysis of Production of Bio-Oil from Catalytic Pyrolysis of Olive Mill Wastewater Sludge with Two Different Cooling Mechanisms. Energy Convers. Manag. X 2022, 13, 100170. [Google Scholar] [CrossRef]
  263. Wang, Y.; Akbarzadeh, A.; Chong, L.; Du, J.; Tahir, N.; Awasthi, M.K. Catalytic Pyrolysis of Lignocellulosic Biomass for Bio-Oil Production: A Review. Chemosphere 2022, 297, 134181. [Google Scholar] [CrossRef] [PubMed]
  264. Vasalos, I.A.; Lappas, A.A.; Kopalidou, E.P.; Kalogiannis, K.G. Biomass Catalytic Pyrolysis: Process Design and Economic Analysis. Wiley Interdiscip. Rev. Energy Environ. 2016, 5, 370–383. [Google Scholar] [CrossRef]
  265. Rajesh Banu, J.; Preethi; Kavitha, S.; Gunasekaran, M.; Kumar, G. Microalgae Based Biorefinery Promoting Circular Bioeconomy-Techno Economic and Life-Cycle Analysis. Bioresour. Technol. 2020, 302, 122822. [Google Scholar] [CrossRef]
Figure 1. Major sources of biomass (reproduced from Jawaid et al., 2017) [30].
Figure 1. Major sources of biomass (reproduced from Jawaid et al., 2017) [30].
Energies 16 05547 g001
Figure 2. Biomass compositions (Modified from Zhang et al., 2014; Thakur and Thakur 2015; and Mansor et al., 2019) [33,34,36].
Figure 2. Biomass compositions (Modified from Zhang et al., 2014; Thakur and Thakur 2015; and Mansor et al., 2019) [33,34,36].
Energies 16 05547 g002
Figure 3. Thermochemical conversion processes with main products (regenerated from Bridgwater 2006 and Patel et al., 2016) [55,56].
Figure 3. Thermochemical conversion processes with main products (regenerated from Bridgwater 2006 and Patel et al., 2016) [55,56].
Energies 16 05547 g003
Figure 4. Types of pyrolysis processes (reproduced from Tripathi et al., 2016) [71].
Figure 4. Types of pyrolysis processes (reproduced from Tripathi et al., 2016) [71].
Energies 16 05547 g004
Figure 5. The pyrolysis processes, according to the movement of feedstock (regenerated from Jahirul et al. 2012 and Campuzano et al., 2019) [86,88].
Figure 5. The pyrolysis processes, according to the movement of feedstock (regenerated from Jahirul et al. 2012 and Campuzano et al., 2019) [86,88].
Energies 16 05547 g005
Figure 6. Major products from the pyrolysis process (regenerated from Chen and Lin 2016 and Reza et al., 2023) [15,66].
Figure 6. Major products from the pyrolysis process (regenerated from Chen and Lin 2016 and Reza et al., 2023) [15,66].
Energies 16 05547 g006
Figure 7. Arrangement of the catalytic pyrolysis processes (reproduced from Luo and Resende 2016; Muneer et al. 2019; Zadeh et al., 2020) [175,176,177].
Figure 7. Arrangement of the catalytic pyrolysis processes (reproduced from Luo and Resende 2016; Muneer et al. 2019; Zadeh et al., 2020) [175,176,177].
Energies 16 05547 g007
Table 1. Type of catalysts and pyrolysis setup used for catalytic pyrolysis of biomass.
Table 1. Type of catalysts and pyrolysis setup used for catalytic pyrolysis of biomass.
FeedstocksCatalystsPyrolyzer TypeRef.
Pine sawdustZSM-5Fluidized bed[157]
Rice huskZSM-5Fixed bed[158]
Aspen woodZSM-5Tubular bed[159]
Pinewood chipZSM-5Fixed bed[160]
Pine barksZSM-5Fixed bed[161]
SawdustZSM-5Fluidized bed[162]
BeechwoodZSM-5Fixed bed[163]
Nannochloropsis sp.ZSM-5Fixed bed[164]
CorncobZSM-5Fluidized bed[165]
Pine sawdust oakMordenitePacked bed[166]
BeechwoodFeO, Al2O3Fixed bed[163]
Hazelnut bagasseFeO, Al2O3Fixed bed[167]
Sewage sludgeLa2O3Fluidized bed[168]
Rice huskZnOFixed bed[169]
Rice huskAl-MSU-F, Al-MCM-41, Rice husk ashFixed bed[158]
MicroalgaeNa2CO3Fixed bed[170]
CelluloseK2CO3Fixed bed[171]
ReedK2CO3Fixed bed[172]
Table 2. Effects of a catalyst on the pyrolysis of rice husk [158].
Table 2. Effects of a catalyst on the pyrolysis of rice husk [158].
SampleCatalystProducts Yield/wt%Major Components in Bio-Oil
BiocharSyngasBio-OilOrganic AcidsAromatic HydrocarbonKetonesFuransHeavy PhenolsLight PhenolsANHYDROSUGARS
Rice husk--41.9218.4739.6116.061.5114.758.9913.5712.773.65
ZSM-542.2719.4538.2915.492.2412.969.6313.4814.442.04
Al-MSU-F43.3119.1839.5915.511.5713.829.5714.2412.562.34
Al-MCM-4143.1518.8039.9814.171.4213.659.5714.9912.602.88
BRHA42.2721.6238.2916.161.5314.769.4812.8412.662.75
Table 3. Effects of natural catalysts in pyrolysis.
Table 3. Effects of natural catalysts in pyrolysis.
SampleCatalystTemperature, °CProducts, wt%Reference
BiocharBio-OilSyngas
Tulip tree Sand 45017.9549.0321.84[203]
Dolomite45026.0434.2639.70
Table 4. Catalyst effects of activated carbon on the pyrolysis process [213].
Table 4. Catalyst effects of activated carbon on the pyrolysis process [213].
SampleCatalystProducts, wt%Major Components, wt%Ref.
Bio-oilSyngasBiocharPhenolsGuaiacolsFuransKetones/Aldehydes
Douglas fir sawdust pelletsNo catalyst45.2011.8043.002.54048.783.9010.36[213]
GAC 830 PLUS (Bituminous coal)31.0044.6024.4025.7415.6514.2416.85
DARCO MRX (Wood)26.5052.6720.8371.871.33-3.03
DARCO 830 (Lignite coal)28.9747.5323.574.77--2.02
Catalyst/Biomass PhenolLevoglucosenone2-Furancarboxaldehyde, 5-methyl-
Glucose063.318.817.6-29.96.8[211]
0.2852.818.724.04.718.88.8
0.449.518.925.610.920.510.8
0.745.818.829.134.5-24.4
1.1344.618.729.688.4--
Table 5. Effects of metal oxide as a catalyst on the pyrolysis process.
Table 5. Effects of metal oxide as a catalyst on the pyrolysis process.
SampleCatalystProducts, wt%Major Components, wt%Ref.
Bio-oilBiocharSyngasAcetic AcidAnhydrosugarFuranPhenol
Palm empty fruit bunches No catalyst42.327.230.51.950.220.381.54[223]
CaO, 5%39.929.430.71.530.070.391.17
CaO, 10%40.430.728.90.90.10.380.64
MgO, 5%39.32931.62.320.060.491.11
MgO, 10%42.325.7322.080.040.460.85
ZnO, 5%44.72728.22.290.070.411.2
ZnO, 10%42.226.8312.230.10.521.2
Isochrysis MicroalgaeNo catalyst16.8237.8645.32 [222]
20.1333.9645.91
18.0730.7651.17
CeO220.8234.8344.35
22.9731.0845.95
19.9128.4851.61
TiO222.4133.1844.41
24.3029.5346.17
21.1026.3952.51
Al2O319.3636.4744.17
22.1034.3443.56
20.0331.8048.17
Table 6. Catalytic Pyrolysis of Acid-washed Cladophora glomerata [228].
Table 6. Catalytic Pyrolysis of Acid-washed Cladophora glomerata [228].
SampleCatalystProducts Yield (wt%)
BiocharBio-OilSyngas
Acid-washed Cladophora Glomerata (ACG)No catalyst344620
Zn/Activated Carbon373528
Table 7. pH of bio-oil for catalysts and non-catalysts [232].
Table 7. pH of bio-oil for catalysts and non-catalysts [232].
SampleCatalystTemperature (°C)pHReference
Cotton stalkNo catalyst6002.8 (±0.1)[232]
CaO4.8 (±0.1)
MgO3.9 (±0.2)
Al2O34.1 (±0.1)
ZSM-53.9 (±0.2)
MCM-412.8 (±0.1)
CuO4.1 (±0.2)
Fe2O33.5 (±0.1)
NiO3.9 (±0.1)
ZnO4.3 (±0.1)
ZrO24.0 (±0.2)
TiO24.0 (±0.1)
Brunei rice huskNo catalyst450 3[158]
ZSM-52.74
Al-MCM-412.83
Al-MSU-F2.69
BRHA2.79
Pinus halepensisSand4502.9[233]
Sand-CaO4.3
Sand-CaO.MgO4.6
Empty fruit bunch (EFB)No catalyst5503.8 ± 0.2[234]
1% CaO3.9 ± 0.2
5% CaO3.92 ± 0.2
10% CaO4.16 ± 0.3
Table 8. Water content in bio-oil for catalytic and nonanalytic pyrolysis [231].
Table 8. Water content in bio-oil for catalytic and nonanalytic pyrolysis [231].
SampleCatalystTemperature (°C)Water (wt%)Reference
Pinewood chipsNo catalyst42520.8 ± 3.9[231]
45021.0 ± 4.6
47520.3 ± 3.9
50020.6 ± 3.9
Brunei rice huskNo catalyst450 (°C)52.60[158].
ZSM-555.56
Al-MCM-4154.66
Al-MSU-F54.64
BRHA55.43
Pinus halepensisSand4503.5[233].
Sand-CaO5.9
Sand-CaO·MgO6.3
Table 9. HHV of bio-oil from the pyrolysis of different feedstocks with various catalyst loading effects [22,180,231].
Table 9. HHV of bio-oil from the pyrolysis of different feedstocks with various catalyst loading effects [22,180,231].
SampleCatalystPrimary Reactor Temperature, °CHHV-Wet Basis, MJ/kgHHV-Dry Basis, MJ/kg
Pinewood chipsNo catalyst42518.6 ± 0.8
45019.1 ± 1.3
47518.4 ± 0.5
50019.7 ± 1.2
Brunei Rice huskNo catalyst450 13.6128.71
ZSM-513.3330.01
Al-MCM-4114.4131.79
Al-MSU-F14.9833.02
BRHA17.1738.53
Empty Fruit Bunch (EFB)No catalyst55013.46 ± 1.98
1% CaO14.95 ± 0.58
5% CaO15.80 ± 0.21
10% CaO17.61 ± 1.94
Table 10. Elements of bio-oil from the pyrolysis of Anchusa azurea [239].
Table 10. Elements of bio-oil from the pyrolysis of Anchusa azurea [239].
Temperature, °CComponentsNo catalystNa2CO3ZnCl2Al2O3Ca(OH)2
350C, wt%45.6856.0046.5249.8250.76
H, wt%7.157.488.196.967.16
N, wt%1.573.371.742.282.64
O, wt%45.6033.1543.5540.9439.44
H/C molar ratio1.871.602.111.671.69
O/C molar ratio0.740.440.700.610.58
550C, wt%45.5957.0647.4149.9057.01
H, wt%7.117.487.217.437.05
N, wt%1.582.712.141.442.89
O, wt%45.7232.7543.2441.2333.05
H/C molar ratio1.871.571.821.781.48
O/C molar ratio0.750.430.680.620.43
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Reza, M.S.; Iskakova, Z.B.; Afroze, S.; Kuterbekov, K.; Kabyshev, A.; Bekmyrza, K.Z.; Kubenova, M.M.; Bakar, M.S.A.; Azad, A.K.; Roy, H.; et al. Influence of Catalyst on the Yield and Quality of Bio-Oil for the Catalytic Pyrolysis of Biomass: A Comprehensive Review. Energies 2023, 16, 5547. https://doi.org/10.3390/en16145547

AMA Style

Reza MS, Iskakova ZB, Afroze S, Kuterbekov K, Kabyshev A, Bekmyrza KZ, Kubenova MM, Bakar MSA, Azad AK, Roy H, et al. Influence of Catalyst on the Yield and Quality of Bio-Oil for the Catalytic Pyrolysis of Biomass: A Comprehensive Review. Energies. 2023; 16(14):5547. https://doi.org/10.3390/en16145547

Chicago/Turabian Style

Reza, Md Sumon, Zhanar Baktybaevna Iskakova, Shammya Afroze, Kairat Kuterbekov, Asset Kabyshev, Kenzhebatyr Zh. Bekmyrza, Marzhan M. Kubenova, Muhammad Saifullah Abu Bakar, Abul K. Azad, Hridoy Roy, and et al. 2023. "Influence of Catalyst on the Yield and Quality of Bio-Oil for the Catalytic Pyrolysis of Biomass: A Comprehensive Review" Energies 16, no. 14: 5547. https://doi.org/10.3390/en16145547

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

Article Metrics

Back to TopTop