[go: up one dir, main page]

0% found this document useful (0 votes)
16 views47 pages

1 s2.0 S174396712300065X AAM

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
16 views47 pages

1 s2.0 S174396712300065X AAM

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 47

Graphical Abstract

1 A critical review on biomass pyrolysis: Reaction mechanisms, process


2 modeling and potential challenges
3
4 Arun Krishna Vuppaladadiyam1*, Sai Sree Varsha Vuppaladadiyam2, Vineet Singh Sikarwar, 3,4,5, Ejaz
5 Ahmad6, Kamal K. Pant7, S. Murugavelh8, Ashish Pandey7, Sankar Bhattacharya2, Ajit Sarmah9,
6 Shao-Yuan Leu1*
7 1
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hong Kong
8 2
Department of Chemical and Biochemical Engineering, Monash University, Clayton, Victoria 3800, Australia.
9 3
Institute of Plasma Physics of the Czech Academy of Sciences, v. v. i., Za Slovankou 1782/3, 18200 Prague 8,
10 Czech Republic.
11 4
Department of Power Engineering, University of Chemistry and Technology Prague, Technická 5, 166 28
12 Prague 6, Czech Republic.
13 5
Department of Green Chemistry and Technology, Ghent University, 9000 Ghent, Belgium.
14 6
Department of Chemical Engineering, Indian Institute of Technology (Indian School of Mines), Dhanbad
15 826004, India
16 7
Department of Chemical Engineering, Indian Institute of Technology Delhi, 110016, India
17 8
CO2 Research and Green Technologies Centre, VIT, Vellore, Tamil Nadu 632014, India.
18 9
Department of Civil and Environmental Engineering, The Faculty of Engineering, The University of Auckland,
19 Private Bag 92019, Auckland 1142, New Zealand
20
21 *Corresponding Author Email:
22 Vuppaladadiyam, A.K., Tel: (+852) 9494 7653. Email: arunk.vuppaladadiyam@polyu.edu.hk; Leu, S.-Y., Tel:
23 (+852) 3400 8322. Email: syleu@polyu.edu.hk, Department of Civil and Environmental Engineering, The Hong
24 Kong Polytechnic University, Hong Kong
25

1
26 Abstract
27 Pyrolysis is a versatile technology for exploiting diversified feedstocks to produce a wide range of
28 products, including biochar, bio-oil, and syngas with high potential in diverse applications. The cardinal
29 motivation of pyrolysis research is to productively use diverse biomass to reduce adverse impacts on ecology
30 and enhance process economics. However, complex reactions of pyrolysis pose operational challenges. Thus,
31 the present review targets the reaction mechanisms and kinetics of pyrolysis to enhance the understanding for
32 better process control, improved performance, and product distribution. Pyrolysis mechanisms of the major
33 structural components of biomass, such as cellulose, lignin, and hemicellulose, as well as proteins, lipids, and
34 carbohydrates, are discussed in detail. Various modeling techniques and tools, viz., mathematical, kinetic,
35 computational fluid dynamic modeling, and machine learning algorithms, have been employed to better
36 understand the pyrolysis mechanisms and product distribution. In addition, the most critical challenges, namely
37 aerosol formation, tar formation and their removal mechanisms, that severely impact the pyrolysis process and
38 products are identified and reported. Thus, the present work critically discusses state-of-art biomass pyrolysis,
39 focusing on the reaction mechanism, modeling, and associated challenges to overcome, given that the pyrolysis
40 products and the process are enhanced.
41
42 keywords: Pyrolysis, Mechanisms, Modeling, Lignocellulosic Biomass, Aquatic Biomass, aerosol and tar
43 formation
44 1. Introduction
45 Clean energy technology banks on biomass owing to its abundance and carbon neutrality (Chio et al., 2019).
46 Biomass can be processed via thermochemical or biochemical conversion routes to generate heat and power.
47 High-value- chemicals derived from biomass have economic benefits and bio-fuels can substitute fossils (Leng
48 et al., 2022). The benefits of employing biomass as a feedstock for energy production are many, including
49 availability in abundance, economical, and being carbon neutral. The estimated global biomass reserves are ca.
50 5 billion tons, making them the fourth most abundantly available resource (Ansari et al., 2021). It is noteworthy
51 that identifying a suitable feedstock is crucial in deciding the conversion route. Structural composition classifies
52 biomass into lignocellulosic and aquatic biomass. Cellulose (37-50%), lignin (15-26%), and hemicellulose (22-
53 32%) dominate lignocellulosic biomass composition (Hameed et al., 2019). Similarly, aquatic biomass is
54 composed of proteins (14-65%), carbohydrates (3-30%), and lipids (1-51%), depending on the specie (Su et al.,
55 2022; Vuppaladadiyam et al., 2018).
56 The biorefinery concept is not new, and the feedstock type delineates biorefineries into four generations. The
57 first-generation biorefinery considered edible crops and animal fats to produce biofuels. Agricultural residues,
58 industrial waste, and municipal and household waste are used in second-generation biorefineries to avoid
59 problems their predecessor raises. The third generation considered photosynthetic aquatic organisms, such as
60 micro- and macro-algae, to produce biofuels and other upgraded products. Finally, the fourth-generation
61 biorefinery relies on the genetic modification of microorganisms to produce biofuels (Su et al., 2022). Major
62 biomass-to-energy conversion routes include physical, agrochemical, biochemical, and thermochemical routes
63 (Ansari et al., 2021; Ghodake et al., 2021). Pyrolysis, a thermochemical process, offers excellent control over
64 process parameters, resulting in low emissions of harmful gases. In addition, the scale-up of pyrolysis plants is

2
65 simple and convenient, unlike incineration plants (Uzakov et al., 2018). Although biomass pyrolysis has been
66 extensively studied, it is undeniable that biomass valorization remains challenging, given the complex nature of
67 biomass. In the last few decades, enormous interest has been in understanding pyrolysis mechanisms and
68 modeling owing to its potential for commercialization and profitability. In addition, the availability of inherent
69 metals in the biomass can profoundly impact the pyrolysis products. Investigating the catalytic impact on the
70 mechanism is vital in reactor design and process optimization. Despite the availability of biomass mechanism in
71 the literature, a majority is focusing on biomass as a whole and detailed mechanism on the individual
72 components is sparse and not comprehensive.
73 In view of process optimization, modeling and simulation of biomass pyrolysis are extremely important. The
74 models usually consider the reaction of individual structural components proceeding independently; thus,
75 biomass modeling is the superposition of these components. The inherent interactions between the structural
76 components are often neglected (Hameed et al., 2019). Modeling of biomass pyrolysis does not include the
77 influence of inorganic biomass species, resulting in deviations from the actual mechanism (Trendewicz et al.,
78 2015). Recent reviews (Anca-Couce, 2016; Sharma et al., 2015) presented an excellent analysis of the
79 conversion rate, reactor models and mechanisms of biomass pyrolysis at molecular, particle and reactor levels.
80 The current work complements the available literature and critically discusses key topics not addressed in the
81 previous reviews. The role of structural components of biomass and secondary reactions on product formation
82 mechanisms has been critically discussed and summarised. The review is structured in four sections; section 1
83 offers basic information on pyrolysis, reaction mechanism and modeling. Section 2 provides an in-depth
84 discussion of the pyrolysis mechanism considering the type/nature of biomass. The overall reaction scheme and
85 each component's reaction mechanism have been summarised. Section 3 provides a detailed review of the
86 modeling of biomass pyrolysis, considering different modeling approaches. Section 4 presents a detailed
87 overview and in-depth discussion of the major challenges in biomass pyrolysis. Aerosol formation mechanisms
88 and their impact on bio-oil, tar formation and removal mechanisms and their impact on pyrolysis have been
89 critically summarised. Finally, section 5 provides information on the prospects in research and development.
90 2. Pyrolysis mechanisms
91 Understanding the biomass pyrolysis mechanisms improves reactor design and augments process
92 optimization. The thermal behavior of the macromolecular components in lignocellulosic and aquatic biomass
93 overlaps with one another. For instance, the pyrolysis of hemicellulose and carbohydrates occurs at 150-200°C,
94 whereas the decomposition of cellulose and proteins occurs at 200-350°C. Similarly, the pyrolytic
95 decomposition of lignin and lipids occurs in the temperature range of 200-450°C (Van de Velden et al., 2010).
96 Char formation results in an aromatic polycyclic structured solid residue (McGrath et al., 2003). Intramolecular
97 and intermolecular reorganizations favor char formation pathways and result in a high degree of thermal
98 stability of the residual char (Fu et al., 2010). Polycyclic structure in char involves the formation and orientation
99 of benzene with the evolution of water and non-condensable gases (NCGs) (Van de Velden et al., 2010).
100 Depolymerization involves in cleavage of bonds among the monomers of the biopolymers. The degree of
101 polymerization decreases until the newly formed molecules are converted into volatiles (Mamleev et al., 2009).
102 However, the condensable molecules are liquid-derived-monomers, dimers or trimers at ambient temperature
103 (Mullen and Boateng, 2011). Fragmentation involves breaking the covalent bonds in the monomer units,

3
104 producing NCGs and a wide range of small-chain organic compounds that are condensable at room temperature
105 (Collard and Blin, 2014). Pyrolysis temperature is a critical parameter that influences the yield and quality of the
106 desired product during pyrolysis. Generally, the pyrolysis temperature is chosen based on the desired product or
107 application. At low pyrolysis temperatures (<500°C), the yield of solid char is high, while at higher
108 temperatures (>700°C), the yield of char decreases. The char quality is also influenced by the pyrolysis
109 temperature, with higher temperatures resulting in higher ash content and lower fixed carbon content. The yield
110 and composition of liquid bio-oil strongly depend on the pyrolysis temperature. Low-temperature pyrolysis
111 (<500°C) typically results in high yields of bio-oil, while high-temperature pyrolysis (>700°C) produces less
112 bio-oil but with higher energy content and lower oxygen content. The bio-oil produced at higher temperatures
113 also has a higher heating value and is more stable. The pyrolysis temperature also influences the composition of
114 gaseous products such as hydrogen, methane, and carbon monoxide. The pathways involved in the primary
115 conversion mechanisms of biomass constituents are presented in Fig. 1. The unstable, volatile compounds of
116 primary reactions may enter secondary reactions, which include recondensation and cracking, and produce low
117 molecular weight (LMW) and high molecular weight (HMW) molecules. During cracking, LMW molecules are
118 formed due to the breakdown of chemical bonds in volatile compounds (Neves et al., 2011). Due to the
119 possibility of breaking a similar chemical bond in the volatiles or the polymer, the products obtained from
120 fragmentation and cracking reactions share similarities. High molecular weight (HMW) molecules are formed in
121 recondensation due to the combination of volatile compounds (Hosoya et al., 2007). Secondary char formation
122 can be attributed to inter-pore recombination (Neves et al., 2011; Wei et al., 2006). In addition, the secondary
123 mechanism can be catalyzed at the reactor surface (Collard and Blin, 2014).

124
125 Fig. 1. Important routes in the primary mechanisms during the conversion of biomass (MW: molecular weight;
126 M: monomer). Adapted from (Collard and Blin, 2014).
127
128 3. Conversion mechanisms for constituents of biomass during pyrolysis
129 The primary reaction involves heating the biopolymers, and releasing volatile compounds through lysis and
130 reorientation of chemical bonds (Hosoya et al., 2007; Van de Velden et al., 2010). These volatiles are relatively
131 unstable and may undergo supplementary conversion via secondary reactions. The breakdown of the chemical

4
132 bonds follows three pathways: char formation, depolymerization and disintegration (Collard and Blin, 2014;
133 Elyounssi et al., 2012). Understanding biomass characteristics is crucial to defining pyrolysis parameters and the
134 final by-product characteristics and distribution. Fig. 2 presents the composition of different types of biomasses:
135 agricultural residues (rice husk, wheat straw, swine manure), forestry biomass (bamboo, beech wood),
136 municipal solid waste (sewage sludge) and industrial residues (sawdust, sugarcane bagasse and tires). The
137 critical moisture content of biomass is <10% for effective thermochemical conversion. Sewage sludge (SS)
138 often has a moisture content of ca. 80% depending on the previous treatment processes and requires a drying
139 step before pyrolysis, which can significantly increase the cost of the overall process. While materials like nut
140 shells, spruce wood, tires and wheat straw have high fixed carbon, typically above 20%, SS and animal manure
141 are low in fixed carbon and high in ash content due to high levels of inorganic elements such as phosphorus,
142 silicon and magnesium that are present in urine and feces (Magdziarz et al., 2016). Other metals, such as
143 aluminum and iron, arise from metal salts used in wastewater treatment plants to precipitate phosphorus and
144 remove nitrogen. Nutshells have the highest lignin content, while the cellulose and hemicellulose percentages
145 are similar. By contrast, SS and animal manure displayed a high amount of cellulose (>50%), a smaller amount
146 of hemicellulose (~12%), and a lower lignin content than 8%.
147 3.1. Pyrolysis mechanism for individual components in lignocellulosic biomass
148 The composition of various lignocellulosic biomasses and other categories of wastes is shown in Fig. 2
149 and Fig. 3(a) summarises the cellulose pyrolysis. A set of reactions (R1-R4) are expected to happen during the
150 two-step cellulose pyrolysis process. After the depolymerization reaction (R1) and formation of intermediate
151 liquid compounds (ILC), competition arises between the fragmentation reaction (R2) and transglycosilation
152 reaction (R3). The products from reactions R3 and R4 are mainly LMWCs and levoglucosan (LGA),
153 respectively.(Anca-Couce, 2016) The products from primary pyrolysis may experience other secondary
154 reactions to produce char. Char production via reaction R4 ishighly debatable (Lin et al., 2009). The pathways
155 associated with the charring reactions (R5) result in the formation of H2O, CO2 and volatile PAHs. The increase
156 in the volatiles retention time and pressure, low temperatures and the availability of inorganics favor secondary
157 charring reactions (Yang et al., 2007), while reaction R6, which is the cracking of volatile compounds in the gas
158 phase, results in CO production at temperatures above 500°C. Recently, a novel thin-film pyrolysis technique
159 was used to reveal the pathways of cellulose pyrolysis. Furan formation occurred directly from cellulose and no
160 intermediates (such as glucose) were detected. Glycosidic linkages undergo homolytic cleavage to form furan
161 and glycolaldehyde against ionic mechanisms. The study also emphasized the interconnection between intra-
162 pyran chemistry and the cleavage of glycosidic bonds (Mettler et al., 2012a). The literature regarding the
163 hemicellulose pyrolysis mechanism is limited, and xylan, a straight-chain polymer of xylose, is an auxiliary to
164 hemicellulose to explain the reaction mechanism of hemicellulose (Gao et al., 2020; Giudicianni et al., 2019;
165 Yang, H. et al., 2020; Zheng et al., 2019). Many remarkable differences exist along with considerable
166 similarities between cellulose and hemicellulose pyrolysis. For instance, hemicellulose pyrolysis starts
167 comparatively at low temperatures to yield more char and a lower quantity of sugars than cellulose (Balci et al.,
168 1993; Orfão et al., 1999; Shafizadeh et al., 1972).
169 Patwardhan et al. (Patwardhan et al., 2011) demonstrated the difference between cellulose and
170 hemicellulose pyrolysis reaction pathways (Fig. 3 (a). Alkali conditions increase the formation of char and CO2

5
171 and decrease the yield of sugars, there is no noticeable influence on the yield of CO. Acetic acid, 1-hydroxy-2-
172 propanone, furfural, CO2, CO and H2O are a few important decomposition products from xylan (Wang, Shurong
173 et al., 2013). O-acetyl groups connected to the primary xylan chain dissociates at low temperature to produce
174 acetic acid (Peng and Wu, 2010). Lignin has a much-complicated structure, and its pyrolysis occurs in several
175 stages and over a wide range of temperatures (Vuppaladadiyam, 2019a; 2019b; 2019c). Initially, lignin softens
176 between 150-190°C, followed by dehydration at ca. 200°C. The cleavage of aliphatic side chains, α- and β-aryl-
177 alkyl ether linkages, methoxyl groups and C-C linkages occur between 150-300, 300, 370-400 and 310-350°C,
178 respectively (Zhou et al., 2014).
179 Lignin pyrolysis produces phenolics, carbonyls, alcohols, water vapor and gases such as CO2 and CO.
180 Methanol and methane are also the typical products of lignin pyrolysis. They are generated in high quantities
181 due to the scission of methoxyl groups. Volatile compounds of lignin pyrolysis are mainly released for two
182 reasons; (a) instability in the propyl chains and (b) instability within the linkages between monomer units and
183 aromatic rings (Collard and Blin, 2014). The release of volatiles is followed by char formation reactions to form
184 polycyclic aromatic structures (Liu et al., 2008). However, benzene rings are highly stable under an inert
185 atmosphere and tend to accumulate during the reaction (Mu et al., 2013). Inorganic biomass compounds also
186 catalyze the pyrolysis of lignin and influence the composition of pyrolysis products and mass loss rate. Higher
187 char formation of softwood at higher temperatures has been linked to the higher carbon content of softwood
188 lignin (Anca-Couce, 2016). The distinctive polymer structure of lignin resists thermal degradation. Lignin,
189 unlike cellulose, is not a linear polymer with repeating subunits but is composed of a wide range of monolignols
190 or chemically distinct subunits. The most common monolignols are p-hydroxyphenyl (H), guaiacyl (G) and
191 syringal (S) monolignols. These monolignols create radicals due to the existence of extracellular laccases or
192 peroxides and the radicals then form a wide variety of bonds (carbon-carbon or ether) with the growing lignin
193 polymer to form a complex, branched network (Weng et al., 2008).

6
194
195 Fig. 2. Composition of widely reported lignocellulosic biomass and other waste feedstocks (Aerts, 1997; Antal et al., 2000; Backreedy et al., 2005; Bhuiyan et al., 2018; Black et al.,
196 2013; Demirbaş, 2005; Di Blasi et al., 2010; Dyjakon and Noszczyk, 2020; Garcı̀a-Pèrez et al., 2002; Gulyurtlu et al., 2004; He et al., 2019; Lynd et al., 1999; M. Ebeling and M. Jenkins, 1985;
197 Pattanayak et al., 2020; Qu et al., 2011; Rabemanolontsoa and Saka, 2013; Sajdak et al., 2015; Singh et al., 2020; Taherzadeh et al., 1997; Uzun et al., 2010; Vuppaladadiyam, Arun K et al.,
198 2019a; Vuppaladadiyam, Arun K. et al., 2019; Vuppaladadiyam, Arun K et al., 2019c; Ward et al., 2014; Yang et al., 2018; Yazdani et al., 2019; Zeng et al., 2015a; Zhao et al., 2018).
7
199
200 Fig. 3. Schematics of (a) cellulose and (b) lignin pyrolysis
201
202 A simple reaction pathway suggested by Zhao et al. (Zhou et al., 2014) for lignin pyrolysis is presented
203 in Fig. 3(b). Pyrolytic lignin is produced along with char and light condensable gases during primary pyrolysis,
204 as shown in Fig. 3 (b). Post this, pyrolytic lignin further reacts to produce phenolic oligomers, phenolic
205 monomers, char, and light condensable gases Kotake et al.(Kotake et al., 2014).
206 3.2. Pyrolysis mechanism for individual components in aquatic biomass
207 The biochemical composition of aquatic biomass, such as micro- and macroalgae, is complex, with
208 proteins, lipids, and carbohydrates as the three primary constituents. Each component contributes, albeit
209 unevenly, to the product formation. The composition of different algal species is presented in Fig. 4. The lipid
210 content in macroalgae is generally lower than in microalgae, varying between 0.1-11.5 and 1.4-51, respectively.
211 The volatile content of macroalgae reached a maximum of 77%, while the microalgae have a maximum of
212 89.4% on a dry ash-free basis. While microalgae seem promising for biogas and bio-oil production, macroalgae
213 appear suitable for biochar production owing to the lower volatile content (low carbon and hydrogen content in
214 general) and high ash content. No trend is noticed regarding the nitrogen, oxygen and sulfur content, and the
215 percentage of these three elements seems to be a characteristic of each algae species. However, apart from two
216 of the species listed in Fig. 4, algae's nitrogen content is usually less than 2.5%. Many research studies
217 considering model compounds have examined the reaction mechanisms and pathways followed by the primary
218 compounds during pyrolysis (Bach and Chen, 2017a, b; Chiaramonti et al., 2017; Du et al., 2013; Wang et al.,
219 2017b). For instance, Wang et al.(Wang et al., 2017a) explored the pyrolysis pathways of microalgae
220 Nannochloropsis sp. and explained that lipids adopted decarbonylation, decarboxylation and fragmentation.
221 While proteins followed deamination, decarboxylation, dimerization and fragmentation, carbohydrates followed
222 dehydrated and fragmentation reactions as major pathways during pyrolysis. Yang et al. (Yang et al., 2019)
223 compiled and summarized possible reaction mechanism pathways presented in Fig. 5.

8
224
225 Fig. 4. Composition of widely reported aquatic biomass feedstock (Álvarez-Viñas et al., 2019; Alves et al., 2019; Andrade et al., 2018; Andrade et al., 2020; Azizi et al., 2020; Chen et
226 al., 2014; Choi et al., 2019; Gong et al., 2013; Huang et al., 2017; Ibrahim et al., 2020; Jung et al., 2019; Kim et al., 2012; Kuo et al., 2020; Lee et al., 2014; Li et al., 2017; Li et al., 2016; Li, Y.
227 et al., 2020; López-González et al., 2014; Ma et al., 2020; Maddi et al., 2011; Mokhtar and Munajat, 2019; Norouzi et al., 2016; Norouzi et al., 2017; Pan et al., 2010; Parsa et al., 2018; Parsa et
228 al., 2019; Rahman, 2020; Raikova et al., 2017; Rizzo et al., 2013; Shin et al., 2011; Shuping et al., 2010; Söyler et al., 2017; Verma et al., 2017; Wang, S. et al., 2018a; Wang, S. et al., 2018b;
229 Wang et al., 2019; Wang, Shuang et al., 2013; Xu et al., 2020; Yadavalli et al., 2014; Yang, C. et al., 2020; Yang et al., 2019; Yuan et al., 2019; Zhong et al., 2020).

9
230 Aquatic biomass comprises amino acids, lipids, chlorophyll, glucose, and xylose. The lipids of the
231 biomass form olefins at a higher temperature. Amino acids at higher temperatures enter into the cyclization
232 reaction to form derivatives of pyrazole and indole (Fig. 6). Calvin cycle enhances the aquatic biomass to store
233 carbohydrates in the form of glucose isomers. Glucose undergoes thermal degradation between 180 -220 °C to
234 produce hexane (C6H14) molecules. Xylose is an aldopentose that represents hemicellulose. The cracking of
235 hemicellulose at higher temperatures results in furfurals. The furans present in pyro-oil are actually from the
236 cracking of hemicellulose. In the algal cells, carbohydrates can be found in poly- and oligosaccharides. Glucose
237 is the commonly available monomer in most species of algae (Templeton et al., 2012). However, cellulose,
238 pectin and hemicellulose are also reported in the cell membrane of some microalgae, such as green algae. Other
239 forms, such as agar, alginic acid (C6H8O6)n, carrageenan, fucans, laminarin (C6H10O5)n, lipopolysaccharides,
240 mannitol (C6H14O6), and peptidoglycan are also seen in algae biomass (Chen et al., 2013; Dawczynski et al.,
241 2007). The carbohydrate pyrolysis is governed by hydrolysis, cracking and dehydration, producing
242 anhydrosugars and furfurals as their main products. The thermal behavior of algae carbohydrates and cellulose
243 are not similar. A few studies reported different forms of carbohydrates, such as alginic acid, fucoidan,
244 laminarin and mannitol, all of which adopted different degradation pathways (Anastasakis et al., 2011) and had
245 different weight loss zones (Debiagi et al., 2017). For instance, mannitol and alginic acid exhibited a single
246 major weight loss region during pyrolysis, and fucoidan and laminarin showed two major weight loss regions
247 (Debiagi et al., 2017). Their intrinsic chemical structure influences the response of carbohydrates to
248 temperature. Pyrolysis of carbohydrates involves moisture removal, breaking the glycosidic bond,
249 rearrangement and ring scission resulting in deoxygenated olefins (Easton et al., 2018). Furans, pyrroles,
250 pyrazines (Harman-Ware et al., 2013; Saber et al., 2016) and polyaromatic hydrocarbons (PAHs) (Hong et al.,
251 2017) are produced during the pyrolysis of carbohydrates at temperatures above 300°C. Additionally, at
252 temperatures above 500°C, the decomposition of saccharides results in CO2 (Wang et al., 2007).
253 Lipids, long-chain hydrocarbons, act as structural components and energy storage molecules. They are
254 available in algae as glycolipids, phospholipids, triglycerides and fatty acids (Debiagi et al., 2017). During
255 pyrolysis, these compounds may initially undergo a cracking reaction to acyl chains from the glycerol backbone
256 to produce long-chain fatty acids (LCFAs) (Ahmed et al., 2018). Further, acids, alcohols, aldehydes, ketones
257 and short-chain hydrocarbons (SCHs) (like olefins) may be produced from LCFAs via decarboxylation,
258 decarbonization, deoxygenation, and/or cracking reactions (Yang et al., 2019). Finally, aromatic hydrocarbons
259 may be produced from SCHs via aromatization and cyclization reactions (Maher and Bressler, 2007).
260 Kebelmann et al.(Kebelmann et al., 2013) reported the formation of hydrocarbon compounds such as 1-
261 nonadecene, heneicosane and heptadecane during the pyrolysis of Chlorella vulgaris. Pyrolysis of the protein in
262 algal biomass exhibits multistep and multiphase kinetics (Debiagi et al., 2017). Pyrolysis of protein-rich algal
263 biomass resulted in amides, amines, imidazoles, indoles, nitriles, polyheteroaromatics, pyrazines, pyrazoles,
264 pyridines, and pyrroles (Andrade et al., 2018; Harman-Ware et al., 2013). The formation of cyclic and linear
265 amides are unique. Intramolecular cyclization results in cyclic amides, whereas amino acids react with
266 carboxylic acids to form linear amides. Hydrophobic protein fragments may be produced as the main products at
267 low temperatures. At high temperatures, gaseous products such as C2H6, NO and HCN can be linked to the
268 protein decomposition in the algal biomass (Lee et al., 2020). Gallois et al.(Gallois et al., 2007) reported the

10
269 pyrolysis mechanism for 20 amino acids and Choi et al.(Choi and Ko, 2011) noted that large quantities of N-
270 heterotrophic compounds were generated during the pyrolysis of amino acid monomers. Table 1 summarises the
271 reaction mechanism of different biomass components.

272
273 Fig. 5. Possible pathways for aquatic biomass pyrolysis.
274
275 Table 1. Summary of pyrolysis mechanisms for biomass components.
Component Structure Mechanism
Cellulose linear homopolysaccharide of cellobiose m 150–300 °C: Dehydration, Depolymerization.
onomers 300–390 °C: Depolymerization. 380–800 °C:
Charring process
Hemicelluloses heteropolysaccharides 150–270 °C: Dehydration and breaking of less
stable linkages. 270–350 °C: Depolymerization.
350–800 °C: Charring process

Lignin complex three-dimensional amorphous 150–420 °C: Conversion of the alkyl chains and
polymer rupture of some of the linkages between units
380–800 °C: Conversion of the short
substituents of the aromatic rings and charring
process
Proteins Sequence of amino acids linked together to Cracking, Deamination and Cracking+Maillard
form a polypeptide chain

Lipids Heterogeneous group of compounds, Decrboxylation, cracking, Decarbonylation and


mainly composed of hydrocarbon chains Deoxygenation

Carbohydrates Include an aldehyde or ketone group and a Deoxygenation, Decarbonylation, Dehydration,


hydroxyl group Cracking and Cyclization

11
276 4. Mathematical modeling
277 Pyrolysis of biomass and other wastes can be assessed either by experimental means or by developing
278 suitable models. Pyrolysis modeling uses mathematical models to predict the yield and composition of the
279 various products formed during pyrolysis. The models typically incorporate detailed chemical kinetics and
280 thermodynamics to capture the complex reaction pathways and rate dependencies during pyrolysis. Pyrolysis
281 modeling has many applications, including process design and optimization, feedstock selection, and evaluating
282 the environmental and economic impacts of pyrolysis processes. Experimental investigations are crucial for the
283 accurate design and subsequent optimization of pyrolyzers. However, experimental studies are expensive and
284 time-consuming, and under extreme operational conditions, precise measurement becomes unmanageable
285 (Hameed et al., 2019; Sikarwar and Zhao, 2017). “Virtual” experiments using mathematical modeling can
286 overcome such challenges.(Sharifzadeh et al., 2019) In general, the simulation of pyrolysis can be classified into
287 four classes, namely, (i) Thermodynamic modeling (TDM), (ii) Kinetic modeling (KM), (iii) Computational
288 Fluid Dynamics (CFD) modeling and (iv) Artificial Neural Network (ANN) modeling (Di Blasi, 2008; Sikarwar
289 et al., 2016). Significant studies related to the modeling of pyrolysis are shown in Table 2. The simplest model
290 is thermodynamic modeling (TDM), which is based on the supposition that reacting species inside the pyrolyzer
291 develop the product distribution that would be found if they reacted for an infinite time (Sikarwar et al., 2017),
292 assisting in deducing the pyrolysis products (Gan et al., 2018). TDMs demonstrate the thermodynamic limits for
293 defined operational conditions (Sikarwar and Zhao, 2016). However, the drawback of TDM is the considerable
294 deviation from real-life set-ups (Krutof and Hawboldt, 2020). Therefore, there are limited studies regarding the
295 development of thermodynamic models compared to CFD and KT models.
296 The KM approach overcomes the limitations of thermodynamic models to some extent. Kinetic models take
297 into account the hydrodynamics of the reactor coupled with the kinetics of reactions (Gan et al., 2018), which
298 aids in developing insights into reaction mechanisms and kinetics. KM encompasses a diverse range, from
299 single-step to intricate multiparameter models (Várhegyi et al., 1997). As this approach is complicated, a deep
300 and complete understanding of kinetics and reaction rate constants is not achieved. The evaluation of complex
301 multiphase flows experienced by a pyrolyzer can be approximated by applying CFD principles. With its
302 inherent potency, CFD modeling is widely used to assess biomass/waste pyrolysis (Xiong et al., 2017), and
303 employed for almost all types of pyrolyzers, but it remains complicated (Zhong et al., 2019). Artificial Neural
304 Network modeling is relatively new and is similar to machine learning. It is commonly employed to deduce the
305 non-linear connection between the input and output based on estimating random non-linear
306 functions.(Sunphorka et al., 2017) In addition, a mathematical narrative of the process is not needed to develop
307 the ANN model (Zhang et al., 2020). However, ANN modeling is impossible when the data is restricted, and the
308 inherent lack of an underlying physical model can lead to issues with model extrapolation.
309 4.1. Thermodynamic modeling (TDM)
310 Gagliano et al.(Gagliano et al., 2018) developed two equilibrium models for slow pyrolysis (Pyro_2)
311 and intermediate pyrolysis (Pyro_3). The authors generated the thermodynamic models in MATLAB and
312 evaluated the gas composition and LHV of NCGs and bio-oil. Global and equilibrium chemical reactions govern
313 the transformation of biomass to char, gases, and bio-oil, which formed the basis of these models. Hydrogen,
314 methane, carbon dioxide, carbon monoxide, and water mole fractions were estimated, along with the amount of

12
315 bio-oil. The amount of char and the operational temperature was constant for both of these models. An
316 enhancement in the producer gas and LHV was noticed when the operating temperature was increased from 400
317 to 700°C. Plausible reasons can be the rise in combustibles and the increase in H2 due to the thermal cracking of
318 gases and vapors. Comparing both models, the authors found that the Pyro_2 model was more precise than
319 Pyro_3 vis-à-vis bio-oil yield. When the modeling deductions were assessed concerning trial data, it was
320 reported that the estimates were within the permissible limits for diverse pyrolytic conditions. The
321 thermodynamic models can be employed to examine the thermodynamic limitations, which in turn can be used
322 to optimize the pyrolysis process.
323 A 100 kg/h banana (Musa spp.) waste was employed in developing an Aspen plus-based thermodynamic model
324 (Ighalo and Adeniyi, 2019). The authors assessed the product composition vis-à-vis temperature for diverse
325 types of banana wastes, which included pseudo-stem, banana peels, and banana leaves. Interestingly, pseudo-
326 stem contributed more toward gas yield. It was reported that gas yields at the thermodynamic limit were similar
327 for all the feed materials at low temperatures (450 to 500°C). However, at higher temperatures (500 to 650°C), a
328 variation in the yield was noticed, with leaves producing the smallest quantity of gases. It also noted that leaves
329 produced less bio-oil and (quite obviously) a more significant amount of char. The proximate analysis indicated
330 that the pseudo-stem produced a higher fraction of volatiles. Consequently, it generated higher amounts of
331 producer gas and oil. Banana pseudo-stem was appropriate for oil generation, whereas banana leaves and peel
332 were suitable for char production.
333 In another investigation, physical and chemical characteristics coupled with vapor-liquid phase
334 properties of fast pyrolysis bio-oil were assessed (Krutof and Hawboldt, 2020). Softwood residues-derived fast
335 pyrolysis bio-oil generated using a pilot-scale auger reactor at 450°C entered an advanced distillation curve
336 equipment under 5 kPa. The thermodynamic state points and vapor phase composition data were deduced using
337 an advanced distillation curve apparatus. VMGSim™ was used to develop a bio-oil model employing the
338 UNIQUAC equations of states. The authors deduced the vapor-liquid equilibrium and physicochemical
339 characteristics using a blend of H2O, pyrolytic lignin, inert solids, and multiple organic components. The results
340 were then obtained using a reduced-pressure version of Windom and Bruno's advanced distillation curve (V-
341 ADC) (Hameed et al., 2019). V-ADC enhanced the distillable fraction of bio-oil from 55 vol% (with ambient
342 conditions) to 72 vol%. It was done by restricting the polymerization at elevated temperatures. Thermodynamic
343 models of bio-oil enhance insights into bio-oil bulk and fraction characteristics, promoting its usage in multiple
344 end applications. Lee et al. (Lee et al., 2007) combined thermodynamic and kinetic simulations to evaluate the
345 product distribution obtained from palm oil waste pyrolysis. The authors employed the HSC Chemistry software
346 to provide a thermodynamic approach, whereas the Sandia PSR code was used for kinetic calculations. In
347 addition, the gas compositions were assessed by HSC computations which were then fed into the PSR code to
348 provide a more precise distribution of gaseous products. The authors reported a decrease in char with a rise in
349 gaseous species when the temperature was enhanced. By adding PSC code to a thermodynamic model, the
350 authors could predict C2H2, C2H4, C2H6 and C3H8, which are formed at the expense of char and CH4, mostly
351 between 200 to 700°C. Among the newly generated species, C2H2 formed the largest fraction with 12% and
352 C2H4 as 4.5% at 400°C. The conjunction of the thermodynamic and kinetic approach can enhance the precision
353 of gas composition values and aid in computing other gaseous species found in smaller fractions.

13
354

14
355 Table 2. A selection of pyrolysis modelling studies.
356
Modeling Tool Approach Line of study Reference
type
TD MATLAB Developing and solving non-linear equations Composition and LHV of gases and amount of bio-oil (Gagliano et al., 2018)

TD ASPEN - Product composition w.r.t. temperature (Ighalo and Adeniyi,


PLUS 2019)
TD VMGSim™ UNIQUAC equations of states Physico-chemical and vapor-liquid phase properties of bio-oil (Krutof and Hawboldt,
2020)
TD and - HSC Chemistry computation code and Sandia Product composition and distribution (Lee et al., 2007)
KT PSR code
TD and - Flynn Wall Ozawa (FWO) model and Feasibility of rice hull pyrolysis using limestone and eggshell (Gan et al., 2018)
KT Distributed activation energy Model (DAEM) catalysts
KT - Modelling study followed by experimental Potential of glycerol pyrolysis to produce hydrogen (Fantozzi et al., 2016)
work for validation
KT - DAEM approach, model-free approach, and Assessment of the influence of vulcanization on the fixed bed (Liu et al., 2018)
model-based method pyrolysis of rubber
KT - DAEM Co-pyrolysis of chest-nutshell and polystyrene (Özsin and Pütün, 2018)
KT - Sequential and coupling method (Coats- Relationship of activation energy (Eα) with conversion rate for (Wang et al., 2016)
Redfern method and Kissinger method) cellulose, hemicellulose and lignin pyrolysis
CFD FLUENT multi-component and multi-step reaction Evaluation of pyrolytic product distribution, temperature (Ding et al., 2020)
model and Euler-Euler approach distribution and flow regime in the reactor bed along with the
effect of temperature, superficial gas velocity and bed height
CFD - Eulerian-Eulerian approach Examination of heat transfer characteristics and hydrodynamics (Hooshdaran et al., 2017)
of conical spouted bed pyrolyzer
CFD - Chemical kinetics along with multiphase flow Development of a design tool to scale-up a fast pyrolysis plant (Jalalifar et al., 2020)
dynamics were employed for auger reactor
with the inclusion of rotating reference frame
CFD - Lumped kinetics approach with multi fluid Evaluation of synergistic impact of intra-particle heat (Zhong et al., 2019)
model conduction and particle shrinkage on biomass fast pyrolysis
ANN - Levenberg-Marquardt pathway Fuel production from plastic waste (Abnisa et al., 2019)
ANN - Least Squares Support Vector Machine (LS- Production of biochar from cattle manure (Cao et al., 2016)
SVM)
ANN MATLAB NFTOOL Hydrogen generation from the pyrolysis of wastes such as olive (Karaci et al., 2016)
husk, cotton cocoon shell and tea waste
ANN - - Evaluation of relation between biomass components and kinetic (Sunphorka et al., 2017)
parameters such as activation energy, reaction order and pre-
exponential factor in pyrolysis
15
357 4.2. Kinetic (KT) modeling
358 Fantozzi et al.(2016) investigated the possibility of glycerol as a feedstock for pyrolysis to generate H2,
359 which could either be employed for CHP or as a fuel in the transportation sector. The gaseous yield and H 2
360 fraction can be enhanced by applying optimal operating conditions. A higher yield of NCGs(70 wt%) and
361 enhanced process efficacy between 750 - 800°C. The H2 mole fraction ranged between 44 to 48 vol%. CFD
362 simulation provides more profound insights into glycerol pyrolysis. Gan et al.(Gan et al., 2018) assessed the
363 kinetic and thermodynamic parameters for rice hull pyrolysis with limestone and eggshells. The FWO model
364 (Ozawa, 1992) and the DAEM (McCown and Harrison, 1982) were adopted, with variable temperatures ranging
365 from 50°C to 900°C. The FWO model considers the first-order reactions and is a single-step reaction model.
366 The DAEM considers the devolatilization of biomass, along with numerous first-order reactions. An average Eα
367 for non-catalytic pyrolysis was noticed as 175.4 - 177.7 kJ/mol. Limestone as a catalyst reported decreased Eα of
368 123.3 – 132.5 kJ/mol, which reduced to 96.1 - 100.4 kJ/mol when eggshell was used as a catalyst. The FWO
369 model considers the first-order reactions and is a single-step reaction model. The average Ea for non-catalytic
370 pyrolysis ranges between 175.4 - 177.7 kJ/mol. Some researchers (Liu et al., 2018) conducted in-depth kinetic
371 modeling investigations to evaluate the impact of vulcanization on the fixed bed pyrolysis of natural-,
372 butadiene- and styrene butadiene-rubber. The distributed activation energy approach predictions were more
373 accurate than the model-based approach, suggesting that the cardinal rubber degradation followed a chain
374 reaction. The highest bio-oil generation was 90.82% for natural rubber at 430°C, whereas 90.61% for butadiene
375 rubber and 92.80% for styrene-butadiene rubber at 470°C. The vulcanized rubber, however, displayed a
376 different trend, where sulfur was released at lower temperatures.
377 On the other hand, gaseous species were increased at the expense of sulfur-containing oils at elevated
378 temperatures. Liu et al.(Liu et al., 2019) employed Rhus Typhina biomass to explore the Eα and frequency
379 factors of the pyrolysis process with the aid of the Friedman method. Thermal degradation rates were deduced
380 via a three pseudo-component DAEM, where the pseudo-components denoted hemicellulose, cellulose, lignin
381 and others respectively. A nonlinear dynamic optimization model was coupled with the DAEM to enhance the
382 model's precision. Pseudo-components DAEM accurately deduced the biomass transformation and degradation
383 rate for 10, 20, and 30 K/min. The experimental data and model deductions were in line. Özsin and Pütün(Özsin
384 and Pütün, 2018) examined the co-pyrolysis behavior of chestnut shell and polystyrene and compared it with the
385 individual performances of both feedstocks. In order to carry out the kinetic investigation, a DAEM was
386 developed. The authors reported that the individual activation energies of polystyrene and chestnut shells as
387 208.9 kJ/mol and 175.2 kJ/mol, respectively. In contrast, the Eα for co-pyrolytic degradation was 191.6 kJ/mol.
388 The predictions via models agreed with the findings from Friedman and FWO iso-conversional methods.
389 Wang et al.(Wang et al., 2016) evaluated the kinetic models employing sequential and coupling
390 methods and assessed the pyrolysis of cellulose, lignin, and hemicellulose, taking microcrystalline cellulose,
391 beechwood xylan and organosolv lignin as the model compounds. The relationship of Eα to conversion rate was
392 examined by applying the Isoconversional approach.(Kujirai and Akahira 1925) A one-step reaction model was
393 used for cellulose, whereas a three-stage model for hemicellulose and a two-stage model for lignin was
394 employed. The Eα for cellulose pyrolysis remained unaltered under different conversion rates, indicating that the
395 complete process followed a common reaction pathway. The concept of a typical reaction pathway paved the

16
396 way for applying a single global reaction model to represent the devolatilization during cellulose pyrolysis. The
397 Coats-Redfern method (Coats and Redfern, 1964) and Kissinger (Elder, 1985) pathways were adopted to
398 produce the reaction model. Avrami-Erofeev nucleation model (Allnatt and Jacobs, 1968) explained cellulose
399 pyrolysis with an Eα of 119.2 kJ/mol. On the other hand, a reaction order model explained the pyrolysis of
400 hemicellulose, with the Eα of 93.6 kJ/mol in the first parallel reaction route and lignin with Eα of 114.6 kJ/mol.
401 The authors further demonstrated that cellulose pyrolysis is comparatively more straightforward vis-à-vis
402 hemicellulose and lignin pyrolysis, with parallel and successive reactions in the latter two, modeling their
403 pyrolysis very complicated. It can be deduced from the literature that while kinetic modeling is an essential tool
404 to enhance the understanding of the pyrolysis of different biomasses, incorporating transport mechanisms can
405 generate improved computational tools which can aid in the design and optimization of pyrolyzers. More
406 research is however needed to formulate and validate all-inclusive models.
407 4.3. Computational Fluid Dynamics (CFD) Modeling
408 The biomass pyrolysis process is highly complicated, with thousands of chemical reactions (Sikarwar
409 and Zhao, 2017). Therefore, it is impossible to consider all the reactions while developing a CFD model. In real-
410 life scenarios, appropriate and simplified reaction kinetics is required to depict feedstock devolatilization
411 followed by tar cracking (Xiong et al., 2018), and is crucial as the chemical reactions directly impact the
412 pyrolyzer’s efficacy. In addition, any model should be computationally affordable. Many chemical reactions,
413 reactants and products are usually considered in all the reactor scale CFD models76 and are known as lumped
414 global kinetics. However, comparatively complicated kinetics have also been employed in recent investigations
415 to depict biomass pyrolysis. Ding et al.(Ding et al., 2020) developed a CFD model for fluidized bed pyrolysis of
416 solid waste to evaluate product distribution, temperature distribution and flow regime in the reactor bed and
417 assessed the impact of temperature, superficial gas velocity and bed height. Pyrolysis was defined by multi-
418 component and multi-step reaction models, whereas multiphase flow was described by the Euler-Euler
419 approach. The feedstock was fed at a rate of 5 kg/h. User-defined functions in Fluent software characterized
420 heterogeneous reactions. It was reported that the feedstock has a trivial impact on gas temperature, which
421 became stable at 37.4 s.
422 Moreover, the mass flow rate of products fluctuated with pyrolysis time within ±10% of the average
423 values after achieving a steady fluidized state. In addition, it was also noticed that there was a periodic
424 fluctuation in the distribution of products in the bed. As the temperature increased, the tar fraction increased at
425 temperatures up to 500°C and then started reducing, whereas the gas fraction showed the opposite tendency. It
426 was noticed that secondary tar cracking was elevated above 500°C, leading to a reduction in tar amounts.
427 Superficial gas velocity harmed gas fraction, reducing the tar residence time and decreasing tar cracking. The
428 other impact of increased gas velocity was an enhancement in the product flow rate fluctuation, although the
429 char fraction was not influenced. The fluctuations in flow rates of pyrolytic products were also reported with
430 increasing bed height, with a negligible impact on product fractions. The authors found the temperature
431 distribution was similar, whereas product yields varied concerning experimental validation. Lu et al. (Lu et al.,
432 2021) from NREL investigated the influence of biomass size and composition on fast pyrolysis using the CFD
433 model and step pyrolysis kinetics. The authors validated the model with two sets of experimental data. In
434 addition, the influence of biomass components on the yields of pyrolysis products was investigated using

17
435 sensitivity analysis. The authors also investigated the effect of particle size on heat transfer, hydrodynamics and
436 product yields using a fluidized bed model by coupling chemical reactions with the CFD reactor model. It was
437 noticed that the impact of particle size on hydrodynamic behavior was similar in the reactor. With the progress
438 in pyrolysis, the biomass particle's density fell to 80-90% of their original density and the lighter particles began
439 to shift to the upper region of the bed. In addition, it was noticed that the larger particles needed longer
440 residence time because they have a slower heating rate and, thus, slow conversion. The simulation also indicated
441 that particle size significantly influenced the yields of pyrolysis products. An increase in bio-oil yield was
442 attributed to the extended residence time of the larger particle size.
443 Hoosdaran et al.(Hooshdaran et al., 2017) evaluated a conical spouted bed pyrolyzer's heat transfer
444 characteristics and hydrodynamics by generating a CFD model employing an Eulerian-Eulerian technique
445 combined with the kinetic theory of granular flow. The study assessed the heat transfer coefficients, pressure
446 drop, particle velocity and solid volume fraction along with the influence of the specularity coefficient (φ), and
447 particle-wall restitution coefficient on heat transfer and hydrodynamics. The authors found a direct relationship
448 between wall temperature, heat transfer coefficients, and pressure drop. However, the opposite trend was
449 noticed for the influence of φ on the bed pressure drop and wall-to-bed heat transfer coefficients. At lower φ, the
450 rate of fall for pressure-drop and heat transfer coefficient was higher, where the rate of decrease was more
451 important for the former than for the latter. The authors deduced that the pressure drop changed by 14%, while
452 the alteration in heat transfer coefficient was about 7% when φ was elevated from 0.01 to 0.15. In addition, an
453 inverse relationship between the restitution coefficient and pressure drop was established. A similar trend was
454 noticed for heat transfer performance. A decrease of 31% in pressure drop and 7.8% in heat transfer coefficient
455 was deduced when the restitution coefficient was increased by 80%. Jalalifar et al.(Jalalifar et al., 2020)
456 developed a CFD model for pilot scale fast pyrolysis intending to use it as a design tool for scale-up. Chemical
457 kinetics and multiphase flow dynamics were employed for an auger reactor, where a rotating reference frame
458 was included to model the impact of the rotation of the auger. Feedstock flow rates varied from 1 to 4 kg/h, the
459 temperature ranged from 400 to 600°C and pressure varied from 0 to 50 kPa. In addition, the impact of N2 was
460 evaluated as a carrier gas with a flow rate ranging from 1 to 10 kg/h and the influence of the angular velocity of
461 the screw was assessed from 45 to 95 rpm. The study reported 500°C to be the optimal temperature for bio-oil
462 generation. It was also noticed that an elevation in bio-oil happens with the enhancement in feedstock flow rate
463 due to the decrease in vapor residence time which resulted in minimizing the reactions of the non-condensable
464 part in the vapor phase. In addition, an increase in bio-oil fraction was noticed when N2 was fed because of
465 reduced vapor residence time. The optimal value of rpm was reported as 70. The results were in good agreement
466 with experimental values.
467 CFD principles were applied by Zhong et al.(Zhong et al., 2019) to investigate the synergistic influence
468 of particle shrinkage and intra-particle heat conduction on the fast pyrolysis of biomass. A lumped kinetics
469 approach was adopted to assess fast pyrolysis, whereas a multi-fluid model was chosen to evaluate the
470 hydrodynamics. Four different cases with diverse particle shrinkage and intra-particle heat conduction
471 arrangements were considered and their efficacies vis-à-vis product yields were investigated. The authors
472 assessed the char characteristics, product distribution and the distribution of particle diameter and density. They
473 reported a decrease in tar fraction and an increase in char amounts when the impact of particle shrinkage and

18
474 intra-particle heat conduction were considered. In addition, a trivial effect of intra-particle heat conduction was
475 noticed, especially for small diameter (325 µm), which further weakened when the particle shrinkage effect was
476 considered. The survey of CFD modeling studies emphatically reflects a need to improve modeling precision.
477 Regarding multi-fluid modeling, the impact of sub-grid structures should be considered (Xiong et al., 2018;
478 Xiong et al., 2017). Moreover, the modeling speed needs to be enhanced for discrete phase modeling as it is
479 computationally intensive. More developments in applying CFD principles to develop deep insights can lead to
480 a better understanding of the pyrolysis process.
481 4.4. Artificial Neural Network (ANN) Modeling
482 Artificial Neural Networks (ANNs) are a programming concept built based on biological neural
483 networks in the human brain. A simple schematic of ANN with two hidden layers forward propagation network
484 is shown in Fig. 6. An ANN comprises nodes called artificial neurons. Depending on the complexity of the
485 model, neurons are organized in multiple layers, which are an input layer, an output layer and several hidden
486 layers (two hidden layers as shown in Fig. 6). ANNs are a widely accepted powerful tool to simulate highly non-
487 linear processes. However, many researchers highlight the lack of interpretability and the black-box nature of
488 ANNs as major drawbacks (Li, J. et al., 2020; Pandey et al., 2016; Serrano and Castelló, 2020). It is to be noted
489 that the researchers have made substantial efforts to address these issues and improve their interpretability
490 (Ascher et al., 2022). The potential of ANN to model highly non-linear problems enhanced its reputation in
491 modeling complex thermochemical processes. ANNs are highlighted as universal function approximators, which
492 means that a neural network can be created to model any given function successfully [103].

493
494 Fig. 6. An ANN model with two hidden layers. I and O represent perceptron's inputs and outputs, respectively.
495
496 Cao et al. (Cao et al., 2016) explored the possibility of obtaining biochar through the pyrolysis of cattle
497 manure. The authors combined Least Squares Support Vector Machine (LS-SVM) with the conventional ANN
498 model to perform the investigation. Thirty-three datasets from the lab-scale reactor were taken to develop the
499 model and predict the results. During the exploration, the value of R2 was found to be 0.96 and 0.80,
500 respectively. The authors argued that the LS-SVM model was more precise and robust vis-à-vis the
501 conventional Neural Network Model. These studies can pave the way for further research employing similar
19
502 approaches for other types of waste. Non-recycled plastic waste is a growing problem in almost all nations.
503 Using pyrolysis to transform plastic waste into usable fuel was evaluated by some researchers by developing the
504 ANN model. These researchers assessed the data about fuel generation from plastic waste through the
505 Levenberg-Marquardt pathway (Abnisa et al., 2019). The lowest mean square error (MSE) formed the basis to
506 designate the optimum number of concealed neurons. In order to ensure precision and reliability, statistical
507 evaluation and graphical presentation were employed to examine the model. The optimal solution aims to report
508 the best fit of MSE between output (from the dataset, product yield) and feed-forward neural network (FANN)
509 prediction. The drawback of the FANN system is the limitation in restricting the training and validation errors.
510 Early stopping method can be used to overcome the over-fit or under-fit issues with FANN, thereby improving
511 performance of FANN. The regression analysis revealed that the values for training and validation datasets were
512 above 0.9. In addition, the mean square error values for both datasets were negligible (2.6419 × 10-4 and
513 0.1114). The values of tar, oil and gas acquired from the developed ANN model were examined via a ternary
514 graph and in close agreement with the published literature. This clearly shows that the pyrolytic products can be
515 precisely and reliably evaluated from the ANN modeling approach.
516 Karaci et al.[106] developed a rigorous ANN model to examine hydrogen generation from the
517 pyrolysis of wastes such as olive husk, cotton cocoon shell and tea waste. The authors employed the Neural Net
518 Fitting tool in MATLAB to perform the computations and made the model more rigorous by including a wide
519 number of parameters: quantity of biomass, biomass diversity, and different types of catalysts (ZnCl 2, NaCO3
520 and K2CO3). The study reported a trivial deviation in hydrogen ratio when the mean square error was
521 considered. The authors noticed that the optimal model was produced for olive husk waste at 700°C with 10%
522 ZnCl2. It has been posited that such ANN models can save a huge amount of time needed for experimental work
523 and reduce labor. In addition, they can be a crucial aid for waste managers, planners, and research workers. Care
524 must always be taken when extrapolating beyond the “trained” regimes of such models, though, in fairness, they
525 can sometimes elucidate pathways not considered by human workers. Thus, for example, AlphaGo, a computer-
526 based program that integrates deep neural networks with an advanced search tree, can be considered an example
527 of the potential “intuitive leaps” possible using artificial intelligence (AI). During the initial stages, AlphaGo
528 was made to play against itself many times and could learn from its mistakes. Over time, AlphaGo became
529 stronger and better at learning and decision-making. However, two moves, move 37 and move 78, are
530 recognized as pointers to how AI can go beyond current thinking and influence major aspects of the world, such
531 as health care and software designs. In a game between AlphaGo and Lee Sedol (the world champion) in March
532 2016, the machine made a move, move 37, which goes beyond what was achieved by human intuition in the
533 3000 years of history of the Chinese game GO. In response, Lee Sedol’s move 78 was described as “God’s
534 touch” by the GO gaming community. In other words, the machine-human interaction during the game
535 empowered the human brain to realize a part of previously unnoticed reality (DeepMind, 2020). Not just
536 empowering the human brain, AI provides an opportunity to enhance the coding process further to provide a
537 better experience from simple tasks (games) to highly complicated and complex tasks (health care).
538 Some researchers (Zhang et al., 2020) developed an ANN model to deduce higher heating value (HV)
539 of producer gas generated via sludge pyrolysis. Multiple parameters were evaluated, such as sludge type, the
540 moisture content in sludge, pyrolytic temperature, catalyst type, and catalyst amount. ANN model with a

20
541 regression coefficient and a model F value of 0.9501 and 19.87 reported goodness of fit. It can be inferred that
542 the model best describes the behavior of the neurons in the hidden layer and the initial value of weights (and
543 bias) on the MSE of the model. The training was automatically stopped after 19 epochs when the network gave
544 the minimum MSE. At the optimal conditions, a three-layer topology network was obtained with eight neurons
545 in the input layer, 15 neurons in the hidden layer, and one neuron in the output layer. A correlation coefficient of
546 0.97 and mean square error of 14.62 was found while forecasting the HV of sludge pyrolysis-derived producer
547 gas. The authors found pyrolytic temperature and sludge moisture as the most crucial parameters influencing the
548 HV of producer gas among all the inputs. They reported the maximum value obtained for HV as 1833.5 kJ/m3N
549 at 895°C with the moisture of 45.63 wt%. They compared the developed model with multiple linear and
550 principal component regression and found it more precise when compared to the latter two approaches.
551 ANN models can be applied for estimating the kinetic parameters of biomass pyrolysis with cellulose,
552 hemicellulose and lignin as inputs. The critical step of development includes the diversity of biomass
553 composition and type. TGA, kinetic data can be employed to calculate the input data. Datasets from around 150
554 TGA experiments of different biomass compositions were taken to examine the network. They reported a non-
555 linear relation between biomass constituents and output parameters. They also argued that the relationship
556 between biomass constituents and kinetic parameters could be accurately forecasted with R2 > 0.9. They adopted
557 30 neurons for reaction order, 17 neurons for Eα and 20 neurons for pre-exponential factor post-optimization
558 with a mean standard error of 0.001. They assessed their results via contour plots and determined that the
559 maximum values of all three kinetic parameters are required by cellulose. These plots also reflected the non-
560 linearity and intricacy of the system. Kasmuri et al.(Kasmuri et al., 2019) conducted an interesting analysis to
561 employ pyrolysis for methanol generation to develop a control system using MATLAB. Model reference control
562 in a dynamic study of nonlinear methanol synthesis vis-à-vis process design and control parameters were used to
563 sustain high methanol generation with set point constant values at optimal operational conditions from
564 experimental investigations. They found the reaction temperature, time and N2 flow to be the vital variables that
565 significantly impacted the process. The maximum yield of ethanol was reported to be 3.09 wt.% with a mean
566 squared error of 0.2617. The nonlinearity of regulating input temperature to the linearity of the measured output
567 of bio-methanol yield was accomplished in this study (Yoon et al., 2019).
568 In recent years, ANN models have become an influential tool for evaluating different dimensions of
569 pyrolysis processes. These models can save a considerable amount of labor and time needed for the
570 experimental work. More importantly, they aid in cost-cutting. Based on this literature survey, it is concluded
571 that combining two or more approaches, such as LS-SVM, fuzzy logic, quantum models, etc., with traditional
572 ANN models enhances the accuracy and reliability of the prediction. Therefore, more research is needed on
573 integrating approaches for diverse scenarios. Modeling approaches have been extensively used to understand the
574 biomass pyrolysis process (Ciesielski et al., 2018). Modeling studies are reported for atomic-scale phenomena
575 (Burnham et al., 2015; Krumm et al., 2016) Meso/Particle-scale phenomena (Ciesielski et al., 2017; Di Blasi,
576 2002a), and the reactor scale (Eri et al., 2017; Lee et al., 2017). Atomic-scale models could consider about 25
577 chemical species for modeling biomass pyrolysis. The model developed by Ranzi et al., employed a categorical
578 reactant and product lumping scheme, which included 25 chemical species (Ranzi et al., 2017). The atomistic
579 and sub-atomistic models provide details that assist in understanding the fundamental reactions of biomass

21
580 pyrolysis. The limitation of atomistic and sub-atomistic models is their link to the physical process, time scale
581 and incomplete product speciation. The single-particle models are computationally fast and simple for particle-
582 scale modeling, but their inability to represent anisotropy, the basic characteristic of biomass particles, is their
583 major disadvantage. However, two- and three-dimensional models consider anisotropy and allow the vapor
584 fields within the particles. The reactor-scale modeling, for example, the CFD approach, although there has been
585 substantial progress in this area, even the most sophisticated modeling approaches consider simple global
586 pyrolysis schemes, assume oversimplified biomass particle geometries (in case of a Lagrangian approach) and
587 employs models that were developed for other materials with different properties. Though substantial research
588 has been done using modeling to understand the pyrolysis process, it is still possible to develop kinetic
589 mechanisms that de-couple transport phenomena with reaction kinetics and facilitate their applicability to new
590 systems. In addition, alongside maintaining computational manageability, future work needs to focus on
591 including chemical speciation that enables predicting product composition.
592
593 5. Challenges in the pyrolysis process
594 Pyrolysis technologies are generally tested at laboratory and pilot scales. Scaling up to commercial production
595 can be challenging owing to the complexity of the heat transfer quality and quantity of the feedstock. Biomass
596 feedstocks are regularly heterogeneous in composition with huge differences in moisture content, particle size,
597 and chemical composition leading to non-uniform heating and reaction rates during pyrolysis, affecting the yield
598 and quality of the products. The following section throws light on the challenges associated with thermal
599 cracking.

600 5.1. Aerosol formation during the pyrolysis process


601 Lignocellulosic biomass, with porous and microstructure structural compounds, under pyrolytic
602 conditions produce a short-lived intermediate liquid phase which furthers breakdown to volatiles and permanent
603 gases and releases aerosols. The inevitable aerosol emissions during pyrolysis can severely impact the
604 downstream processes. The first-ever experiment of biomass pyrolysis done in a micropyrolyzer revealed that
605 the feed, before final decomposition, is converted into a short-lived intermediate (liquid phase) accompanied by
606 multiple pyrolysis reactions (Ansari et al., 2021). Aerosols formed during the pyrolysis process can be grouped
607 as primary or secondary. Three mechanisms are responsible for the formation of primary aerosols during the
608 pyrolysis process; (i) intermediate liquid (IL) mechanism, in which biomass forms an IL during the thermal
609 decomposition and within the liquid phase, reactions such as dehydration, depolymerization and rearrangement
610 occur before the products are converted to gas, char or aerosols (Teixeira et al., 2016). (ii) Vapor-bubble
611 collapse mechanism, within which the molten IL vapor-bubbles are produced, is followed by a collapse to form
612 a liquid jet and droplets of liquid aerosol (Iisa et al., 2019). (iii) Film-aerosol generation occurs due to the shear
613 thinning of bubbles, and the aerosols formed via this mechanism generate smaller aerosol droplets compared to
614 the vapor-bubble collapse mechanism (Teixeira et al., 2016). However, it is possible to trap the aerosols inside a
615 solid matrix after ejection and, as a consequence of the collision, the particle size of the aerosols may increase
616 (Pecha et al., 2019). Secondary aerosols can be formed due to secondary condensation reactions during the rapid
617 cooling of pyrolysis vapors (Winkelmann et al., 2018).

22
618 Even though secondary aerosols have been reported for an extended time, the existence of primary
619 aerosols in biomass pyrolysis is recently reported. Very few studies characterized the size distribution of
620 primary aerosols at the laboratory scale. Consequently, little is known about aerosols' composition, size
621 distribution and capacity to transport the impurities (such as inorganics) from biomass to the final products (bio-
622 oil) (Teixeira et al., 2016). Teixeira et al.(Teixeira et al., 2011) proposed a ‘reactive boiling ejection’
623 mechanism, which suggests that the primary aerosols are spontaneously generated from biomass (cellulose)
624 pyrolysis. A high-speed photography technique was used to record the release of primary aerosols from molten
625 cellulose. The aerosols were released in the form of liquid jet as the bubble collapsed, as shown in Fig. 7. The
626 authors reported that after 100 ms, particles are fully molten and adopt a hemispherical shape (Fig. 7B).
627 Spontaneous ejection of aerosol occurs (107 ms), and a trace of the ejected particle is visible at 108 ms (Fig.
628 7C). The ejected particle appears to slow down after one ms (Fig. 7D) and the molten droplet is converted to
629 gases, aerosols and vapors, leaving a clean surface (Fig. 4 E). The aerosol droplets were noticed to have a
630 number-based mode diameter of ca. 0.8 mm and a size smaller than 3 µm. In addition, the authors reported that,
631 based on the aerosol composition, the aerosols were formed via the ejection mechanism and not because of
632 secondary reactions.

633
634 Fig. 7. Stages of ejection of aerosols from microcrystalline cellulose particle. Source: Adapted from
635 (Teixeira et al., 2011).
636
637 In another study aimed at understanding the impact of aerosols on the origin and fate of inorganic
638 particles based on product fractionation, the authors reported that ca. 3% of initial feed was transported to the
639 gas phase as primary aerosols (Teixeira et al., 2016). Fig. 8 shows the high-speed imaging of cellulose pyrolysis
640 on a 500 ⁰C aluminum nitride surface. Fig. 8-A shows the sequential frames explaining the solid cellulose
641 particle (0 ms) forming a liquid particle/surface interaction (1183 ms), followed by a four-phase stage (1350 ms)
642 which has a solid cellulose cap (@1, 1350 ms), solid/liquid interface (@2. 1350 ms), liquid intermediate at the
643 bottom of the particle (@3, 1350 ms) and reflection (@4, 1350 ms). Fig. 8-B shows more than 40 large primary
644 aerosols, highlighted in red with varying intensity, ejected from the liquid intermediate cellulose in 1 ms. It is
23
645 reported that ca. 30% of bio-oil originates from particulates and aerosols (Radlein et al., 1987) and more than 60
646 wt.% of inorganic content in the bio-oil is linked to aerosols (Jendoubi et al., 2011).
647 The boiling ejection mechanism can serve as a tool to address the challenge of aerosol generation.
648 However, there could be numerous other possible mechanisms and the information on whether these alternative
649 mechanisms occur in the IL phase is unknown.(Mettler et al., 2012b) Soot is an aerosol product produced due to
650 the incomplete combustion of biomass. Unfortunately, the soot formation mechanisms from biomass fuel are not
651 well understood either (Li et al., 2018). During biomass pyrolysis, macro-molecule cracking results in HMW
652 compounds, such as saccharides, phenols, ketones and aromatics. It is worth stressing that the pathway from
653 these HMW compounds to nascent soot production is very complex and more challenging to explain than that
654 arising from small molecules.(Wang, X. et al., 2018) Considering that high levels of volatiles are produced
655 during biomass pyrolysis, the secondary reactions of hydrocarbons in volatiles are expected to trigger soot
656 formation. There are three main routes for soot formation, namely (i) small hydrocarbon mechanism based on
657 hydrogen abstraction carbon addition (HACA), (ii) generation of aromatic rings via dimerization of two
658 resonance stabilized cyclopentadienyl radicals and (iii) direct condensation and transformation of aromatic rings
659 (Fletcher et al., 1997; Richter and Howard, 2000). However, it is still unknown as to which pathway is dominant
660 in biomass pyrolysis.

661
662 Fig. 8. High-speed imaging of cellulose pyrolyzing on a 500 ⁰C aluminium nitride surface. (A) Sequence
663 indicating the transformation of solid particle to liquid, bubble and vapor. (B) Aerosols observed during one ms
664 within the detection limits. Scale bars = 100 µm.
665
666 5.2. Tar formation and removal
667 Tar is used as a collective term for all the organic compounds with a molecular weight greater than
668 benzene but does not include char and soot. Currently, there are numerous definitions of tar available in the
669 literature depending on the field of research, however, no commonly accepted definition is agreed on (Gredinger
670 et al., 2018). According to standard DIN CEN/TS 15439(GIS, 2013), tar is a “generic (unspecific) term for
24
671 entity of all organic compounds present in the producer gas excluding gaseous hydrocarbons (C1 to C6)”. The
672 total quantity and composition of tar can be determined using a solid phase adsorption method to perform tar
673 measurement. In general, tar is classified into five categories, namely (Sikarwar et al., 2016), (i) GC
674 undetectable tars, (ii) heterocyclic components (for example, C6H6O, C7H8O, etc.), (iii) Aromatic compounds
675 (C7H8, C8H8, etc.), (iv) Light poly-aromatic hydrocarbons (C10H8, C14H10, etc.) and (v) Heavy poly-aromatic
676 hydrocarbons (C16H10, C24H12, etc.).
677 In another classification of tar vis-à-vis pyrolyzers and gasifiers, researchers (Evans and Milne, 1997) have
678 categorized them into the four groups below:
679 i. Primary tars - these originate from the cellulose, hemicellulose and lignin fractions in the biomass.
680 ii. Secondary tars - originate from the conversion of primary tars and mostly contain phenolics and
681 olefins.
682 iii. Alkyl tertiary tars - cardinally contain methyl derivatives of aromatics such as xylene.
683 iv. Condensed tertiary tars - poly-aromatic hydrocarbons.
684 The stability of the bio-oil is severely affected by the polymerization reaction of tar components at
685 ambient conditions. Tar can chemically bind with the porous activated carbon irreversibly, thereby preventing
686 desorption. The above challenges necessitate the removal of tar from the pyro-oil. Several pathways have been
687 suggested as a result of extensive research in the field of tar cracking. These methods are widely employed and
688 are partially effective. In general, tar destruction can be classified into five strategies: self-modification methods,
689 mechanism methods, thermal cracking, plasma cracking and catalytic cracking (Sikarwar et al., 2016).
690
691 5.2.1. Self-modification method
692 As the name suggests, this route involves altering operational variables such as temperature, pressure,
693 heating rate, residence time, etc., to modify the tar produced during the thermochemical process. Sadakata et
694 al.(Sadakata et al., 1987) conducted a study to evaluate the pyrolysis of wood-derived lignin and holocellulose
695 in an electric furnace in the range of 400 to 900 ⁰C with a heating rate of more than 1000°C/min. The authors
696 reported the production of ca. 10% of tar from holocellulose at 400°C, which fell rapidly, with further increase
697 in temperature, to ca. 1% at 700°C. In addition, the maximum amount of tar was seen at 500°C for lignin,
698 followed by a decrease when the temperature was further raised. The authors concluded that pyrolysis of lignin
699 and holocellulose should be conducted above 700°C to ensure the minimum tar yield. In another investigation
700 conducted by Fagbemi et al.(Fagbemi et al., 2001), it was observed that temperature has a crucial impact on tar
701 destruction. Tar quantities were found to increase up to 600°C and then reduce with further increase in
702 temperature, which suggested the existence and enhancement of secondary tar reactions beyond 600 ⁰C as the
703 underlying reason for tar reduction.
704
705 5.2.2. Mechanism method
706 Equipment such as electrostatic precipitators (ESP), scrubbers, cyclones and filters are all mechanical
707 devices (Zeng et al., 2015b). They are employed to remove particles and droplets from the product gas.
708 However, diverse experimental investigations effectively captured tar and other impurities (Han and Kim,
709 2008). For the thermochemical treatment of rice husk, a venturi scrubber displayed an efficiency between 51 to

25
710 91% for tar capture.(Hasler et al., 1997) The cardinal drawback of this system was the generation of huge
711 amounts of wastewater which represented a disposal problem. A novel oil-based tar capture system devised by
712 the Energy Research Center of the Netherlands known as OLGA (oil-based gas washer in Dutch)(Boerrigter et
713 al., 2005) reported an impressive removal of 99% for phenol and 97% for heterocyclic tars. Other advantages of
714 ESP were the lack of fouling of ESP plates and insensitivity to voltage and residence times. In an interesting
715 study, Hasler et al.(Hasler et al., 1997) noted the potential of a bed filter made up of activated carbon in tar
716 removal. The phenols and high-boiling hydrocarbons were reported to be effectively captured by the carbon
717 filter, which was placed near the fabric filter. However, major challenges include cleaning tar from the filter
718 surface and the eventual plugging of such systems.
719
720 5.2.3. Thermal cracking
721 The application of heat to crack tars is called thermal cracking. Di Blasi (Di Blasi, 2002b) studied the
722 effect of feedstock particle size on tar generation and varied the heating rate from 177 to 182°C/s for a reactor
723 temperature of 527°C. The particle size ranged from 0.1 to 6 mm, and woody biomass was the feedstock. The
724 intra-particle reactions of tar cracking did not occur when small particles were used, effectively offering zero
725 residence time of tars within the particles to convert volatiles. Consequently, there was a generation of high
726 amounts of tar (~80%). On the other hand, when large particles of size 5 mm were examined at the temperature
727 of 527°C, high degradation of tars was noticed. Brandt et al. (Brandt and Henriksen, 2000) used a pure
728 aluminum oxide reactor to investigate the impact of temperature on tar cracking. Out of the four sets of
729 measurements, one set was related to the gas from pyrolysis units and one set was from the updraft gasifier. The
730 residence time for these experiments was 0.5 s. Here, a significant tar degradation (130 mg/kg dry feedstock as
731 compared to the initial value of more than 1000 mg/kg dry feedstock) occurred for a residence time of 0.5 s at a
732 temperature of 1250°C or higher. Gas from the updraft gasifier was found to have a tar content of 32 mg/kg dry
733 feedstock, after treatment at 1290°C, compared to the previous value of more than 1000 mg/kg dry feedstock.
734 These studies certainly reflect the significance of particle size and how heat is used to crack tar.
735
736 5.2.4. Plasma cracking
737 Corona discharges were employed in some investigations to assess their influence on tar cracking. Nair
738 et al. (Nair et al., 2004) applied corona discharge for about 3 min to decompose naphthalene. At 400°C with an
739 energy density of 40 J/L, 50% of the naphthalene was destroyed. Tar cracking was due to reactive species
740 generated by high-energy electrons (Sikarwar et al., 2020). The authors argued that an elevation in gas
741 temperature would enhance the oxidation kinetics coupled with primary O radical yield. Heesch et al.(Van
742 Heesch et al., 2000) evaluated the reliability of a plasma system for a gasifier (with wood as the feedstock).
743 They demonstrated a removal efficiency of 72 to 95% for dust, 68% for heavy tar and 50% for light tar. In a
744 study by Pemen et al.(Pemen et al., 2002), gliding arc technology was examined for tar destruction with variable
745 energy density. The temperature range applied for the experimental work was from 400 to 800°C. The study
746 reported a slight enhancement in tar destruction when the energy density was raised. However, a significant and
747 continuous tar decomposition was noticed with rising temperature in the reactor.
748

26
749 5.2.5. Catalytic cracking
750 Catalytic tar decomposition has become one of the well-studied pathways for tar cracking in recent years.
751 Diverse range of catalysts (Ni, Olivine, Dolomite, Zeolites, Ceramic, Carbon-based, etc.) have been
752 investigated. Catalytic tar cracking occurs in both in-situ and secondary tar cracking (Gil et al., 1999). Under in-
753 situ tar cracking, a catalyst is employed inside the pyrolyzer/reactor. In contrast, the secondary tar cracking
754 method uses a secondary reactor, where tar in the producer gas is treated in the presence of a catalyst (Sikarwar
755 et al., 2016). The tar decomposition reactions are often kinetically limited; therefore, the reaction rate is
756 improved by increasing the temperature or using a catalyst. However, a catalyst can enhance the reaction rate
757 only where a reaction is thermodynamically favorable and becomes increasingly unimportant at higher
758 temperatures (Zeng et al., 2020). All the significant reactions normally considered in tar degradation are
759 described in Table 3. Simell et al.(Simell et al., 1997) conducted an in-depth study, used toluene as the model
760 compound for tar and suggested a series of tar cracking reactions (R1 to R8) and equilibrium reactions (R9 to
761 R14) (Table 2). Toluene is taken as a generic depiction of tar in all the reactions. The authors suggested that
762 Group VIII metals catalyzed the reforming reactions (R1 and R6).(Rönkkönen et al., 2010) It was inferred that
763 Fe present in ilmenite, coupled with CO2 and H2O in producer gas, catalysed reactions R1 and R6 and shown by
764 the authors that at 830 ⁰C and above, R6 is more favoured than R1 on the grounds of thermodynamics. However,
765 R12, R13 and R14 were favoured below 650 ⁰C. Moreover, R8 is increased at elevated temperatures. In
766 addition, R9 was found to be catalyzed by Fe-based catalysts (Uddin et al., 2008).
767 Table 3. Significant reactions associated with tar cracking (Shen and Yoshikawa, 2013; Simell et al., 1997).
S. No. Chemical reaction Name of the reaction
R1 CnHm* + nH2O → nCO + (n + 0.5m)H2 Steam reforming
R2 CnHm + xH2O → CxHy + qCO + pH2 Steam dealkylation
R3 CnHm → C + CxHy + gas Thermal cracking
R4 CnHm + (2n – (m/2))H2 → nCH4 Hydro cracking
R5 CnHm + xH2 → CxHy + qCH4 Hydro dealkylation
R6 CnHm + nCO2 → 2nCO + 0.5mH2 Dry reforming
R7 CnH2n+2 → Cn-1H2(n-1) + CH4 Cracking
R8 CnH2n+2 → nC + (n+1)H2 Carbon formation
R9 CO + H2O → H2 + CO2 Water gas shift
R10 CO + 3H2 → CH4 + H2O Methanation 1
R11 2H2 + C → CH4 Methanation 2
R12 CO + H2 → H2O + C Water gas 1
R13 CO2 + 2H2 → 2H2O + C Water gas 2
R14 C + CO2 → 2CO Boudouard
768 *CnHm = Tars
769
770 5. Future recommendations
771 Based on the research carried out in this review, several longstanding challenges are associated with
772 biomass pyrolysis. Though cellulose can offer an approximate approach to elucidate the pyrolysis mechanism, it

27
773 is essential to develop model components that represent actual biomass, including cellulose, lignin, and
774 hemicellulose/ proteins, carbohydrates, and lipids. Further investigations on the fundamental mechanisms are
775 necessary to develop a methodology to optimize the pyrolysis process at the industrial level. Even though
776 different reactor designs have been tested to maximize the bio-oil yields, establishing the design on commercial
777 reactors for the pyrolysis of biomass at an industrial scale is challenging due to techno-economic reasons.
778 Further research is necessary for developing a pyrolyzer with features such as being suitable for a wide range of
779 feedstock, high heat transfer efficiency, low dependency on feedstock characteristics (i.e., moisture, particle
780 size) and high processing capacity for commercial application. Molecular studies on pyrolysis chemistry require
781 immediate attention. Pyrans and furan formation during cellulose pyrolysis can be explored to elucidate a
782 plausible mechanism. Determining the properties of intermediate liquids requires faster response (<1 ms)
783 analytics, necessitating techniques capable of collecting and rendering inert to further reaction of intermediate
784 liquids in a short time. Challenges in the scale-up of pyrolysis technology restrict the commercialization of
785 biomass pyrolysis. Technical challenges, such as developing fundamental descriptions for elementary reaction
786 mechanisms that would facilitate the optimization of the process, should be addressed.
787 Popular reaction models consider a component of biomass rather than biomass. Research needs to be
788 focussed on developing a robust model considering whole biomass, including the major structural components
789 and other components (including inorganic species and extractives). The availability of data for model
790 development remains a significant challenge. Developing new models can improve the prediction accuracy and
791 generalization capacity related to the data type. For instance, modeling the process in a particular type of reactor
792 generates highly accurate data when the experimental/training data corresponds to the particular type of reactor.
793 This indicates that the model applies to a narrow range of training data. Efforts are recommended in developing
794 diverse models with low but satisfactory prediction accuracy. Therefore, more efforts are needed to generate
795 valid and legitimate data by considering a wide range of parameters. Future research is recommended to
796 illuminate the black box of ANN algorithms to enhance the interpretability of the results. Improving the
797 interpretability of the ANN models can help to understand the fundamental mechanisms of the pyrolysis process
798 from the network prediction. The solid-state models reported in the literature mainly considered TGA data, so
799 these models cannot adequately fit the short induction period during the pyrolysis process. It is to be noted that
800 only a limited number of studies considered multiple parallel reaction models, and data in this segment is sparse.
801 The kinetic models that use thermodynamic equilibrium equations often neglect the information about heat and
802 mass transfer phenomena. Further research is to be carried out in developing models that include these aspects.
803 A very urgent research direction for developing CFD modeling is improving model accuracy. Also, little is
804 known about aerosol formation and its chemistry during biomass pyrolysis. Most information regarding the
805 formation of primary aerosols is confined to the bubble ejection mechanism. Research must be directed towards
806 identifying the potential alternative aerosol formation mechanisms, which would require detailed information
807 about the properties of intermediate liquid species. In addition, a complete understanding of generation
808 mechanisms and the process through which the aerosols escape the microstructures of biomass is necessary to
809 address the challenge of aerosol generation.
810 Conclusions

28
811 Conversion technologies such as pyrolysis can utilize a wide range of biomass to generate valuable
812 products such as bio-oil, char, and pyrolysis gases. A deeper understanding of the pyrolysis mechanisms and
813 modeling is necessary for its optimization and cost reduction. A detailed description of the mechanism involved
814 in the conversion of polymers constituting the biomass, this review provides a comprehensive overview on the
815 pyrolysis mechanisms of structural components of lignocellulosic as well as aquatic biomass. It can be
816 concluded that each modeling approach's characteristics are unique under different operating conditions and
817 specific materials. At present, the mechanical models are mostly based on model compounds such as protein,
818 lipids and carbohydrates or lignin, cellulose and hemicellulose. However, more attention is needed in
819 developing robust models that consider the inorganics and extractives and represent the whole biomass. A
820 deeper understanding of pyrolysis is necessary to optimize, scale up, and reduce costs. Understanding aerosol
821 chemistry could minimize the impurities in bio-oil. Higher temperatures and longer residence time during
822 thermal cracking can minimize tar formation.
823 Author Contributions
824 Arun K. Vuppaladadiyam, and Shao-Yuan Leu conceived the idea and designed the manuscript; Arun K.
825 Vuppaladadiyam and Sai Sree Varsha wrote the manuscripts and took the contributions of Vineet Singh
826 Sikarwar, S. Murugavelh, Elsa Antunes, Ajit Sarmah and Sankar Bhattacharya, Ejaz Ahmad and Kamal Pant,
827 gave valuable suggestion and critically evaluated the manuscript many times that improved the quality of the
828 manuscript. Arun K. Vuppaladadiyam and Sai Sree Varsha share equal authorship. All the authors have agreed
829 to submit this manuscript to the Journal of Energy Institute.
830 Abbreviations
831 AI artificial intelligence; ANN artificial neural network; CFD computational fluid dynamics; DAEM distributed
832 activation energy model; FANN feed-forward neural network; FWO Flynn Wall Ozawa; HMW high molecular
833 weight; ILC intermediate liquid; KM kinetic modeling; LCFAs long-chain fatty acids; LHV lower heating
834 value; LMW low molecular weight; compounds; LS-SVM least squares support vector machine; MSE mean
835 square error; NCGs non-condensable gases; SS Sewage sludge; TDM thermodynamic modeling;
836 Data availability Not applicable.
837 Funding
838 The authors thank for the financial supports from the Hong Kong Research Grant Council via General Research
839 Fund (RGC/GRF 15212319), and Hong Kong Postdoctoral Fellowship (PDFS2223-5S05, A.K.
840 Vuppaladadiyam).
841 Competing interests
842 The authors have no financial or proprietary interests in any material discussed in this article.
843 References
844 Abnisa, F., Sharuddin, A., Dayana, S., bin Zanil, M.F., Daud, W., Ashri, W.M., Mahlia, I., Meurah,
845 T., 2019. The Yield Prediction of Synthetic Fuel Production from Pyrolysis of Plastic Waste by
846 Levenberg–Marquardt Approach in Feedforward Neural Networks Model. Polymers 11(11), 1853.
847 Aerts, D., 1997. Co-firing switchgrass in a 50 MW pulverized coal boiler, Fuel and Energy Abstracts.
848 p. 331.

29
849 Ahmed, A., Abu Bakar, M.S., Azad, A.K., Sukri, R.S., Phusunti, N., 2018. Intermediate pyrolysis of
850 Acacia cincinnata and Acacia holosericea species for bio-oil and biochar production. Energy
851 Conversion and Management 176, 393-408.
852 Allnatt, A., Jacobs, P., 1968. Theory of nucleation in solid state reactions. Canadian Journal of
853 Chemistry 46(2), 111-116.
854 Álvarez-Viñas, M., Flórez-Fernández, N., Torres, M.D., Domínguez, H., 2019. Successful
855 Approaches for a Red Seaweed Biorefinery. Marine Drugs 17(11), 620.

856 Alves, J.L.F., Da Silva, J.C.G., da Silva Filho, V.F., Alves, R.F., Ahmad, M.S., Ahmad, M.S.,
857 Galdino, W.V.d.A., De Sena, R.F., 2019. Bioenergy potential of red macroalgae Gelidium floridanum
858 by pyrolysis: Evaluation of kinetic triplet and thermodynamics parameters. Bioresource Technology
859 291, 121892.
860 Anastasakis, K., Ross, A., Jones, J., 2011. Pyrolysis behaviour of the main carbohydrates of brown
861 macro-algae. Fuel 90(2), 598-607.
862 Anca-Couce, A., 2016. Reaction mechanisms and multi-scale modelling of lignocellulosic biomass
863 pyrolysis. Progress in Energy and Combustion Science 53, 41-79.
864 Andrade, L., Batista, F., Lira, T., Barrozo, M., Vieira, L., 2018. Characterization and product
865 formation during the catalytic and non-catalytic pyrolysis of the green microalgae Chlamydomonas
866 reinhardtii. Renewable Energy 119, 731-740.
867 Andrade, L.A., Barbosa, J.M., Barrozo, M.A.S., Vieira, L.G.M., 2020. A comparative study of the
868 behavior of Chlamydomonas reinhardtii and Spirulina platensis in solar catalytic pyrolysis.
869 International Journal of Energy Research 44(7), 5397-5411.
870 Ansari, K.B., Kamal, B., Beg, S., Wakeel Khan, M.A., Khan, M.S., Al Mesfer, M.K., Danish, M.,
871 2021. Recent developments in investigating reaction chemistry and transport effects in biomass fast
872 pyrolysis: A review. Renewable and Sustainable Energy Reviews 150, 111454.
873 Antal, M.J., Allen, S.G., Dai, X., Shimizu, B., Tam, M.S., Grønli, M., 2000. Attainment of the
874 Theoretical Yield of Carbon from Biomass. Industrial & Engineering Chemistry Research 39(11),
875 4024-4031.
876 Ascher, S., Watson, I., You, S., 2022. Machine learning methods for modelling the gasification and
877 pyrolysis of biomass and waste. Renewable and Sustainable Energy Reviews 155, 111902.
878 Azizi, K., Keshavarz Moraveji, M., Arregi, A., Amutio, M., Lopez, G., Olazar, M., 2020. On the
879 pyrolysis of different microalgae species in a conical spouted bed reactor: Bio-fuel yields and
880 characterization. Bioresource Technology 311, 123561.
881 Bach, Q.-V., Chen, W.-H., 2017a. A comprehensive study on pyrolysis kinetics of microalgal
882 biomass. Energy Conversion and Management 131, 109-116.

30
883 Bach, Q.-V., Chen, W.-H., 2017b. Pyrolysis characteristics and kinetics of microalgae via
884 thermogravimetric analysis (TGA): A state-of-the-art review. Bioresource Technology 246, 88-100.
885 Backreedy, R.I., Fletcher, L.M., Jones, J.M., Ma, L., Pourkashanian, M., Williams, A., 2005. Co-
886 firing pulverised coal and biomass: a modeling approach. Proceedings of the Combustion Institute
887 30(2), 2955-2964.
888 Balci, S., Dogu, T., Yucel, H., 1993. Pyrolysis kinetics of lignocellulosic materials. Industrial &
889 engineering chemistry research 32(11), 2573-2579.
890 Bhuiyan, A.A., Blicblau, A.S., Islam, A.K.M.S., Naser, J., 2018. A review on thermo-chemical
891 characteristics of coal/biomass co-firing in industrial furnace. Journal of the Energy Institute 91(1), 1-
892 18.
893 Black, S., Szuhánszki, J., Pranzitelli, A., Ma, L., Stanger, P.J., Ingham, D.B., Pourkashanian, M.,
894 2013. Effects of firing coal and biomass under oxy-fuel conditions in a power plant boiler using CFD
895 modelling. Fuel 113, 780-786.
896 Boerrigter, H., Van Paasen, S., Bergman, P., Könemann, J., Emmen, R., Wijnands, A., 2005. OLGA
897 tar removal technology. Energy research Centre of the Netherlands, ECN-C--05-009.
898 Brandt, P., Henriksen, U.B., 2000. Decomposition of tar in gas from updraft gasifier by thermal
899 cracking, 1st world conference and exhibition on biomass for energy and industry.
900 Burnham, A.K., Zhou, X., Broadbelt, L.J., 2015. Critical Review of the Global Chemical Kinetics of
901 Cellulose Thermal Decomposition. Energy & Fuels 29(5), 2906-2918.
902 Cao, H., Xin, Y., Yuan, Q., 2016. Prediction of biochar yield from cattle manure pyrolysis via least
903 squares support vector machine intelligent approach. Bioresource Technology 202, 158-164.
904 Chen, C.-Y., Zhao, X.-Q., Yen, H.-W., Ho, S.-H., Cheng, C.-L., Lee, D.-J., Bai, F.-W., Chang, J.-S.,
905 2013. Microalgae-based carbohydrates for biofuel production. Biochemical Engineering Journal 78,
906 1-10.
907 Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y., 2014. Thermal decomposition dynamics and
908 severity of microalgae residues in torrefaction. Bioresource Technology 169, 258-264.
909 Chiaramonti, D., Prussi, M., Buffi, M., Rizzo, A.M., Pari, L., 2017. Review and experimental study
910 on pyrolysis and hydrothermal liquefaction of microalgae for biofuel production. Applied Energy 185,
911 963-972.
912 Chio, C., Sain, M., Qin, W., 2019. Lignin utilization: A review of lignin depolymerization from
913 various aspects. Renewable and Sustainable Energy Reviews 107, 232-249.
914 Choi, J.H., Kim, S.-S., Kim, J., Woo, H.C., 2019. Fast pyrolysis of fermentation residue derived from
915 Saccharina japonica for a hybrid biological and thermal process. Energy 170, 239-249.
916 Choi, S.-S., Ko, J.-E., 2011. Analysis of cyclic pyrolysis products formed from amino acid monomer.
917 Journal of Chromatography A 1218(46), 8443-8455.

31
918 Ciesielski, P.N., Pecha, M.B., Bharadwaj, V.S., Mukarakate, C., Leong, G.J., Kappes, B., Crowley,
919 M.F., Kim, S., Foust, T.D., Nimlos, M.R., 2018. Advancing catalytic fast pyrolysis through integrated
920 multiscale modeling and experimentation: Challenges, progress, and perspectives. WIREs Energy and
921 Environment 7(4), e297.
922 Ciesielski, P.N., Wiggins, G.M., Jakes, J.E., Daw, C.S., 2017. CHAPTER 11 Simulating Biomass
923 Fast Pyrolysis at the Single Particle Scale, Fast Pyrolysis of Biomass: Advances in Science and
924 Technology. The Royal Society of Chemistry, pp. 231-253.
925 Coats, A.W., Redfern, J., 1964. Kinetic parameters from thermogravimetric data. Nature 201(4914),
926 68-69.
927 Collard, F.-X., Blin, J., 2014. A review on pyrolysis of biomass constituents: Mechanisms and
928 composition of the products obtained from the conversion of cellulose, hemicelluloses and lignin.
929 Renewable and Sustainable Energy Reviews 38, 594-608.
930 Dawczynski, C., Schubert, R., Jahreis, G., 2007. Amino acids, fatty acids, and dietary fibre in edible
931 seaweed products. Food chemistry 103(3), 891-899.
932 Debiagi, P.E.A., Trinchera, M., Frassoldati, A., Faravelli, T., Vinu, R., Ranzi, E., 2017. Algae
933 characterization and multistep pyrolysis mechanism. Journal of Analytical and Applied Pyrolysis 128,
934 423-436.
935 DeepMind, 2020. AlphaGo in "https://deepmind.com/research/case-studies/alphago-the-story-so-far".
936 Assessed on 8th December 2020 at 23:00.
937 Demirbaş, A., 2005. Thermochemical Conversion of Biomass to Liquid Products in the Aqueous
938 Medium. Energy Sources 27(13), 1235-1243.
939 Di Blasi, C., 2002a. Modeling intra- and extra-particle processes of wood fast pyrolysis. AIChE
940 journal 48(10), 2386-2397.
941 Di Blasi, C., 2002b. Modeling intra‐and extra‐particle processes of wood fast pyrolysis. AIChE
942 journal 48(10), 2386-2397.
943 Di Blasi, C., 2008. Modeling chemical and physical processes of wood and biomass pyrolysis.
944 Progress in Energy and Combustion Science 34(1), 47-90.
945 Di Blasi, C., Branca, C., Galgano, A., 2010. Biomass Screening for the Production of Furfural via
946 Thermal Decomposition. Industrial & Engineering Chemistry Research 49(6), 2658-2671.
947 Ding, K., Xiong, Q., Zhong, Z., Zhong, D., Zhang, Y., 2020. CFD simulation of combustible solid
948 waste pyrolysis in a fluidized bed reactor. Powder Technology 362, 177-187.
949 Du, Z., Hu, B., Ma, X., Cheng, Y., Liu, Y., Lin, X., Wan, Y., Lei, H., Chen, P., Ruan, R., 2013.
950 Catalytic pyrolysis of microalgae and their three major components: Carbohydrates, proteins, and
951 lipids. Bioresource Technology 130, 777-782.
952 Dyjakon, A., Noszczyk, T.J.E., 2020. Alternative Fuels from Forestry Biomass Residue: Torrefaction
953 Process of Horse Chestnuts, Oak Acorns, and Spruce Cones. Energies 13(10), 2468.
32
954 Easton, M.W., Nash, J.J., Kenttämaa, H.I., 2018. Dehydration Pathways for Glucose and Cellobiose
955 During Fast Pyrolysis. The Journal of Physical Chemistry A 122(41), 8071-8085.
956 Elder, J.P., 1985. The general applicability of the Kissinger equation in thermal analysis. Journal of
957 thermal analysis 30(3), 657-669.
958 Elyounssi, K., Collard, F.-X., Mateke, J.-a.N., Blin, J., 2012. Improvement of charcoal yield by two-
959 step pyrolysis on eucalyptus wood: A thermogravimetric study. Fuel 96, 161-167.
960 Eri, Q., Zhao, X., Ranganathan, P., Gu, S., 2017. Numerical simulations on the effect of potassium on
961 the biomass fast pyrolysis in fluidized bed reactor. Fuel 197, 290-297.
962 Evans, R.J., Milne, T.A., 1997. Chemistry of tar formation and maturation in the thermochemical
963 conversion of biomass, Developments in thermochemical biomass conversion. Springer, pp. 803-816.
964 Fagbemi, L., Khezami, L., Capart, R., 2001. Pyrolysis products from different biomasses: application
965 to the thermal cracking of tar. Applied energy 69(4), 293-306.
966 Fantozzi, F., Frassoldati, A., Bartocci, P., Cinti, G., Quagliarini, F., Bidini, G., Ranzi, E.M., 2016. An
967 experimental and kinetic modeling study of glycerol pyrolysis. Applied Energy 184, 68-76.
968 Fletcher, T.H., Ma, J., Rigby, J.R., Brown, A.L., Webb, B.W., 1997. Soot in coal combustion systems.
969 Progress in Energy and Combustion Science 23(3), 283-301.
970 Fu, P., Hu, S., Xiang, J., Li, P., Huang, D., Jiang, L., Zhang, A., Zhang, J., 2010. FTIR study of
971 pyrolysis products evolving from typical agricultural residues. Journal of Analytical and Applied
972 Pyrolysis 88(2), 117-123.
973 Gagliano, A., Nocera, F., Bruno, M., Blanco, I., 2018. Effectiveness of thermodynamic adaptative
974 equilibrium models for modeling the pyrolysis process. Sustainable Energy Technologies and
975 Assessments 27, 74-82.
976 Gallois, N., Templier, J., Derenne, S., 2007. Pyrolysis-gas chromatography–mass spectrometry of the
977 20 protein amino acids in the presence of TMAH. Journal of Analytical and Applied Pyrolysis 80(1),
978 216-230.
979 Gan, D.K.W., Loy, A.C.M., Chin, B.L.F., Yusup, S., Unrean, P., Rianawati, E., Acda, M.N., 2018.
980 Kinetics and thermodynamic analysis in one-pot pyrolysis of rice hull using renewable calcium oxide
981 based catalysts. Bioresource Technology 265, 180-190.
982 Gao, Z., Li, N., Wang, Y., Niu, W., Yi, W., 2020. Pyrolysis behavior of xylan-based hemicellulose in
983 a fixed bed reactor. Journal of Analytical and Applied Pyrolysis 146, 104772.
984 Garcı̀a-Pèrez, M., Chaala, A., Roy, C., 2002. Vacuum pyrolysis of sugarcane bagasse. Journal of
985 Analytical and Applied Pyrolysis 65(2), 111-136.
986 Ghodake, G.S., Shinde, S.K., Kadam, A.A., Saratale, R.G., Saratale, G.D., Kumar, M., Palem, R.R.,
987 Al-Shwaiman, H.A., Elgorban, A.M., Syed, A., Kim, D.-Y., 2021. Review on biomass feedstocks,
988 pyrolysis mechanism and physicochemical properties of biochar: State-of-the-art framework to speed
989 up vision of circular bioeconomy. Journal of Cleaner Production 297, 126645.
33
990 Gil, J., Caballero, M.A., Martín, J.A., Aznar, M.-P., Corella, J., 1999. Biomass gasification with air in
991 a fluidized bed: effect of the in-bed use of dolomite under different operation conditions. Industrial &
992 Engineering Chemistry Research 38(11), 4226-4235.
993 GIS, 2013. Biomass gasification - Tar and particles in product gases - Sampling and analysis. German
994 Institute for Standardisation (Deutsches Institut für Normung), Germany.
995 Giudicianni, P., Gargiulo, V., Alfè, M., Ragucci, R., Ferreiro, A.I., Rabaçal, M., Costa, M., 2019.
996 Slow pyrolysis of xylan as pentose model compound for hardwood hemicellulose: A study of the
997 catalytic effect of Na ions. Journal of Analytical and Applied Pyrolysis 137, 266-275.
998 Gong, X., Zhang, B., Zhang, Y., Huang, Y., Xu, M., 2013. Investigation on pyrolysis of low lipid
999 microalgae Chlorella vulgaris and Dunaliella salina. Energy & Fuels 28(1), 95-103.
1000 Gredinger, A., Spörl, R., Scheffknecht, G., 2018. Comparison measurements of tar content in
1001 gasification systems between an online method and the tar protocol. Biomass and Bioenergy 111,
1002 301-307.
1003 Gulyurtlu, I., Abelha, P., Gregório, A., García-García, A., Boavida, D., Crujeira, A., Cabrita, I., 2004.
1004 The Emissions of VOCs during Co-Combustion of Coal with Different Waste Materials in a Fluidized
1005 Bed. Energy & Fuels 18(3), 605-610.
1006 Hameed, S., Sharma, A., Pareek, V., Wu, H., Yu, Y., 2019. A review on biomass pyrolysis models:
1007 Kinetic, network and mechanistic models. Biomass and Bioenergy 123, 104-122.
1008 Han, J., Kim, H., 2008. The reduction and control technology of tar during biomass
1009 gasification/pyrolysis: An overview. Renewable and Sustainable Energy Reviews 12(2), 397-416.
1010 Harman-Ware, A.E., Morgan, T., Wilson, M., Crocker, M., Zhang, J., Liu, K., Stork, J., Debolt, S.,
1011 2013. Microalgae as a renewable fuel source: fast pyrolysis of Scenedesmus sp. Renewable energy 60,
1012 625-632.
1013 Hasler, P., Buehler, R., Nussbaumer, T., 1997. Evaluation of Gas Cleaning Technologies for Small
1014 Scale Biomass Gasifiers: Biomass Programme; Final Report. Bundesamt für Energiewirtschaft.
1015 He, Q., Guo, Q., Ding, L., Wei, J., Yu, G., 2019. CO2 gasification of char from raw and torrefied
1016 biomass: Reactivity, kinetics and mechanism analysis. Bioresource Technology 293, 122087.
1017 Hong, Y., Chen, W., Luo, X., Pang, C., Lester, E., Wu, T., 2017. Microwave-enhanced pyrolysis of
1018 macroalgae and microalgae for syngas production. Bioresource Technology 237, 47-56.
1019 Hooshdaran, B., Hosseini, S.H., Haghshenasfard, M., Esfahany, M.N., Olazar, M., 2017. CFD
1020 modeling of heat transfer and hydrodynamics in a draft tube conical spouted bed reactor under
1021 pyrolysis conditions: Impact of wall boundary condition. Applied Thermal Engineering 127, 224-232.
1022 Hosoya, T., Kawamoto, H., Saka, S., 2007. Pyrolysis behaviors of wood and its constituent polymers
1023 at gasification temperature. Journal of Analytical Applied Pyrolysis 78(2), 328-336.

34
1024 Huang, F., Tahmasebi, A., Maliutina, K., Yu, J., 2017. Formation of nitrogen-containing compounds
1025 during microwave pyrolysis of microalgae: Product distribution and reaction pathways. Bioresource
1026 Technology 245, 1067-1074.
1027 Ibrahim, A.F.M., Dandamudi, K.P.R., Deng, S., Lin, J.Y.S., 2020. Pyrolysis of hydrothermal
1028 liquefaction algal biochar for hydrogen production in a membrane reactor. Fuel 265, 116935.
1029 Ighalo, J.O., Adeniyi, A.G., 2019. Thermodynamic modelling and temperature sensitivity analysis of
1030 banana (Musa spp.) waste pyrolysis. SN Applied Sciences 1(9), 1086.
1031 Iisa, K., Johansson, A.-C., Pettersson, E., French, R.J., Orton, K.A., Wiinikka, H., 2019. Chemical
1032 and physical characterization of aerosols from fast pyrolysis of biomass. Journal of Analytical and
1033 Applied Pyrolysis 142, 104606.
1034 Jalalifar, S., Abbassi, R., Garaniya, V., Salehi, F., Papari, S., Hawboldt, K., Strezov, V., 2020. CFD
1035 analysis of fast pyrolysis process in a pilot-scale auger reactor. Fuel 273, 117782.
1036 Jendoubi, N., Broust, F., Commandre, J.M., Mauviel, G., Sardin, M., Lédé, J., 2011. Inorganics
1037 distribution in bio oils and char produced by biomass fast pyrolysis: The key role of aerosols. Journal
1038 of Analytical and Applied Pyrolysis 92(1), 59-67.
1039 Jung, J.-M., Kim, S., Lee, J., Oh, J.I., Choi, Y.-E., Kwon, E.E., 2019. Tailoring pyrogenic products
1040 from pyrolysis of defatted Euglena gracilis using CO2 as reactive gas medium. Energy 174, 184-190.
1041 Karaci, A., Caglar, A., Aydinli, B., Pekol, S., 2016. The pyrolysis process verification of hydrogen
1042 rich gas (H–rG) production by artificial neural network (ANN). International Journal of Hydrogen
1043 Energy 41(8), 4570-4578.
1044 Kasmuri, N., Kamarudin, S., Abdullah, S., Hasan, H., Som, A.M., 2019. Integrated advanced
1045 nonlinear neural network-simulink control system for production of bio-methanol from sugar cane
1046 bagasse via pyrolysis. Energy 168, 261-272.
1047 Kebelmann, K., Hornung, A., Karsten, U., Griffiths, G., 2013. Intermediate pyrolysis and product
1048 identification by TGA and Py-GC/MS of green microalgae and their extracted protein and lipid
1049 components. Biomass and Bioenergy 49, 38-48.
1050 Kim, S.-S., Ly, H.V., Choi, G.-H., Kim, J., Woo, H.C., 2012. Pyrolysis characteristics and kinetics of
1051 the alga Saccharina japonica. Bioresource Technology 123, 445-451.
1052 Kotake, T., Kawamoto, H., Saka, S., 2014. Mechanisms for the formation of monomers and oligomers
1053 during the pyrolysis of a softwood lignin. Journal of Analytical and Applied Pyrolysis 105, 309-316.
1054 Krumm, C., Pfaendtner, J., Dauenhauer, P.J., 2016. Millisecond Pulsed Films Unify the Mechanisms
1055 of Cellulose Fragmentation. Chemistry of Materials 28(9), 3108-3114.
1056 Krutof, A., Hawboldt, K.A., 2020. Thermodynamic model of fast pyrolysis bio-oil advanced
1057 distillation curves. Fuel 261, 116446.
1058 Kujirai , T., Akahira , T., 1925. Effect of temperature on the deterioration of fibrous insulating
1059 materials. Scientific papers of the Institute of Physical and Chemical Research (Tokyo) 2, 223-252.
35
1060 Kuo, P.-C., Illathukandy, B., Wu, W., Chang, J.-S., 2020. Plasma gasification performances of
1061 various raw and torrefied biomass materials using different gasifying agents. Bioresource Technology
1062 314, 123740.
1063 Lee, D.H., Yang, H., Yan, R., Liang, D.T., 2007. Prediction of gaseous products from biomass
1064 pyrolysis through combined kinetic and thermodynamic simulations. Fuel 86(3), 410-417.
1065 Lee, H.W., Choi, S.J., Park, S.H., Jeon, J.-K., Jung, S.-C., Kim, S.C., Park, Y.-K., 2014. Pyrolysis and
1066 co-pyrolysis of Laminaria japonica and polypropylene over mesoporous Al-SBA-15 catalyst.
1067 Nanoscale Research Letters 9(1), 376.
1068 Lee, J.E., Park, H.C., Choi, H.S., 2017. Numerical Study on Fast Pyrolysis of Lignocellulosic
1069 Biomass with Varying Column Size of Bubbling Fluidized Bed. ACS Sustainable Chemistry &
1070 Engineering 5(3), 2196-2204.
1071 Lee, X.J., Ong, H.C., Gan, Y.Y., Chen, W.-H., Mahlia, T.M.I., 2020. State of art review on
1072 conventional and advanced pyrolysis of macroalgae and microalgae for biochar, bio-oil and bio-
1073 syngas production. Energy Conversion and Management 210, 112707.
1074 Leng, E., Guo, Y., Chen, J., Liu, S., E, J., Xue, Y., 2022. A comprehensive review on lignin pyrolysis:
1075 Mechanism, modeling and the effects of inherent metals in biomass. Fuel 309, 122102.
1076 Li, B., Yang, L., Wang, C.-q., Zhang, Q.-p., Liu, Q.-c., Li, Y.-d., Xiao, R., 2017. Adsorption of Cd(II)
1077 from aqueous solutions by rape straw biochar derived from different modification processes.
1078 Chemosphere 175, 332-340.
1079 Li, J., Pan, L., Suvarna, M., Tong, Y.W., Wang, X., 2020. Fuel properties of hydrochar and pyrochar:
1080 Prediction and exploration with machine learning. Applied Energy 269, 115166.
1081 Li, Y., Cui, J., Zhang, G., Liu, Z., Guan, H., Hwang, H., Aker, W.G., Wang, P., 2016. Optimization
1082 study on the hydrogen peroxide pretreatment and production of bioethanol from seaweed Ulva
1083 prolifera biomass. Bioresource Technology 214, 144-149.
1084 Li, Y., Tan, H., Wang, X., Bai, S., Mei, J., You, X., Ruan, R., Yang, F., 2018. Characteristics and
1085 Mechanism of Soot Formation during the Fast Pyrolysis of Biomass in an Entrained Flow Reactor.
1086 Energy & Fuels 32(11), 11477-11488.
1087 Li, Y., Zhu, C., Jiang, J., Yang, Z., Feng, W., Li, L., Guo, Y., Hu, J., 2020. Hydrothermal liquefaction
1088 of macroalgae with in-situ-hydrogen donor formic acid: Effects of process parameters on products
1089 yield and characterizations. Industrial Crops and Products 153, 112513.
1090 Lin, Y.-C., Cho, J., Tompsett, G.A., Westmoreland, P.R., Huber, G.W., 2009. Kinetics and
1091 Mechanism of Cellulose Pyrolysis. The Journal of Physical Chemistry C 113(46), 20097-20107.
1092 Liu, H., Ahmad, M.S., Alhumade, H., Elkamel, A., Cattolica, R.J., 2019. Three pseudo-components
1093 kinetic modeling and nonlinear dynamic optimization of Rhus Typhina pyrolysis with the distributed
1094 activation energy model. Applied Thermal Engineering 157, 113633.

36
1095 Liu, Q., Wang, S., Zheng, Y., Luo, Z., Cen, K., 2008. Mechanism study of wood lignin pyrolysis by
1096 using TG–FTIR analysis. Journal of Analytical and Applied Pyrolysis 82(1), 170-177.
1097 Liu, S., Yu, J., Bikane, K., Chen, T., Ma, C., Wang, B., Sun, L., 2018. Rubber pyrolysis: Kinetic
1098 modeling and vulcanization effects. Energy 155, 215-225.
1099 López-González, D., Fernandez-Lopez, M., Valverde, J., Sanchez-Silva, L., 2014. Pyrolysis of three
1100 different types of microalgae: kinetic and evolved gas analysis. Energy 73, 33-43.
1101 Lu, L., Gao, X., Gel, A., Wiggins, G.M., Crowley, M., Pecha, B., Shahnam, M., Rogers, W.A., Parks,
1102 J., Ciesielski, P.N., 2021. Investigating biomass composition and size effects on fast pyrolysis using
1103 global sensitivity analysis and CFD simulations. Chemical Engineering Journal 421, 127789.
1104 Lynd, L.R., Wyman, C.E., Gerngross, T.U., 1999. Biocommodity Engineering. Biotechnology
1105 Progress 15(5), 777-793.
1106 M. Ebeling, J., M. Jenkins, B., 1985. Physical and Chemical Properties of Biomass Fuels.
1107 Transactions of the ASAE 28(3), 898-0902.
1108 Ma, C., Geng, J., Zhang, D., Ning, X., 2020. Non-catalytic and catalytic pyrolysis of Ulva prolifera
1109 macroalgae for production of quality bio-oil. Journal of the Energy Institute 93(1), 303-311.
1110 Maddi, B., Viamajala, S., Varanasi, S., 2011. Comparative study of pyrolysis of algal biomass from
1111 natural lake blooms with lignocellulosic biomass. Bioresource Technology 102(23), 11018-11026.
1112 Magdziarz, A., Wilk, M., Gajek, M., Nowak-Woźny, D., Kopia, A., Kalemba-Rec, I., Koziński, J.A.,
1113 2016. Properties of ash generated during sewage sludge combustion: A multifaceted analysis. Energy
1114 113, 85-94.
1115 Maher, K.D., Bressler, D.C., 2007. Pyrolysis of triglyceride materials for the production of renewable
1116 fuels and chemicals. Bioresource Technology 98(12), 2351-2368.
1117 Mamleev, V., Bourbigot, S., Le Bras, M., Yvon, J., 2009. The facts and hypotheses relating to the
1118 phenomenological model of cellulose pyrolysis: Interdependence of the steps. Journal of Analytical
1119 Applied Pyrolysis 84(1), 1-17.
1120 McCown, M.S., Harrison, D.P., 1982. Pyrolysis and hydropyrolysis of Louisiana lignite. Fuel 61(11),
1121 1149-1154.
1122 McGrath, T.E., Chan, W.G., Hajaligol, M.R., 2003. Low temperature mechanism for the formation of
1123 polycyclic aromatic hydrocarbons from the pyrolysis of cellulose. Journal of Analytical Applied
1124 Pyrolysis 66(1-2), 51-70.
1125 Mettler, M.S., Mushrif, S.H., Paulsen, A.D., Javadekar, A.D., Vlachos, D.G., Dauenhauer, P.J.,
1126 2012a. Revealing pyrolysis chemistry for biofuels production: Conversion of cellulose to furans and
1127 small oxygenates. Energy & Environmental Science 5(1), 5414-5424.
1128 Mettler, M.S., Vlachos, D.G., Dauenhauer, P.J., 2012b. Top ten fundamental challenges of biomass
1129 pyrolysis for biofuels. Energy & Environmental Science 5(7), 7797-7809.

37
1130 Mokhtar, M., Munajat, N., 2019. Torrefaction of biomass macroalga Ulva intestinalis using TGA, IOP
1131 Conference Series: Earth and Environmental Science: International Conference on Sustainable Energy
1132 and Green Technology. IOP Publishing, Kuala Lumpur, Malaysia, p. 012004.
1133 Mu, W., Ben, H., Ragauskas, A., Deng, Y., 2013. Lignin pyrolysis components and upgrading—
1134 technology review. Bioenergy Research 6(4), 1183-1204.
1135 Mullen, C.A., Boateng, A.A., 2011. Characterization of water insoluble solids isolated from various
1136 biomass fast pyrolysis oils. Journal of Analytical Applied Pyrolysis 90(2), 197-203.
1137 Nair, S., Yan, K., Pemen, A., Winands, G., Van Gompel, F., Van Leuken, H., Van Heesch, E.,
1138 Ptasinski, K., Drinkenburg, A., 2004. A high-temperature pulsed corona plasma system for fuel gas
1139 cleaning. Journal of electrostatics 61(2), 117-127.
1140 Neves, D., Thunman, H., Matos, A., Tarelho, L., Gómez-Barea, A., 2011. Characterization and
1141 prediction of biomass pyrolysis products. Progress in Energy and Combustion Science 37(5), 611-
1142 630.
1143 Norouzi, O., Jafarian, S., Safari, F., Tavasoli, A., Nejati, B., 2016. Promotion of hydrogen-rich gas
1144 and phenolic-rich bio-oil production from green macroalgae Cladophora glomerata via pyrolysis over
1145 its bio-char. Bioresource Technology 219, 643-651.
1146 Norouzi, O., Safari, F., Jafarian, S., Tavasoli, A., Karimi, A., 2017. Hydrothermal gasification
1147 performance of Enteromorpha intestinalis as an algal biomass for hydrogen-rich gas production using
1148 Ru promoted Fe–Ni/γ-Al2O3 nanocatalysts. Energy Conversion and Management 141, 63-71.
1149 Orfão, J.J.M., Antunes, F.J.A., Figueiredo, J.L., 1999. Pyrolysis kinetics of lignocellulosic
1150 materials—three independent reactions model. Fuel 78(3), 349-358.
1151 Ozawa, T., 1992. Estimation of activation energy by isoconversion methods. Thermochimica Acta
1152 203, 159-165.
1153 Özsin, G., Pütün, A.E., 2018. Co-pyrolytic behaviors of biomass and polystyrene: Kinetics,
1154 thermodynamics and evolved gas analysis. Korean Journal of Chemical Engineering 35(2), 428-437.
1155 Pan, P., Hu, C., Yang, W., Li, Y., Dong, L., Zhu, L., Tong, D., Qing, R., Fan, Y., 2010. The direct
1156 pyrolysis and catalytic pyrolysis of Nannochloropsis sp. residue for renewable bio-oils. Bioresource
1157 technology 101(12), 4593-4599.
1158 Pandey, D.S., Das, S., Pan, I., Leahy, J.J., Kwapinski, W., 2016. Artificial neural network based
1159 modelling approach for municipal solid waste gasification in a fluidized bed reactor. Waste
1160 Management 58, 202-213.
1161 Parsa, M., Jalilzadeh, H., Pazoki, M., Ghasemzadeh, R., Abduli, M., 2018. Hydrothermal liquefaction
1162 of Gracilaria gracilis and Cladophora glomerata macro-algae for biocrude production. Bioresource
1163 Technology 250, 26-34.
1164 Parsa, M., Nourani, M., Baghdadi, M., Hosseinzadeh, M., Pejman, M., 2019. Biochars derived from
1165 marine macroalgae as a mesoporous by-product of hydrothermal liquefaction process:
38
1166 Characterization and application in wastewater treatment. Journal of Water Process Engineering 32,
1167 100942.
1168 Pattanayak, S., Hauchhum, L., Loha, C., Sailo, L., 2020. Selection criteria of appropriate bamboo
1169 based biomass for thermochemical conversion process. Biomass Conversion and Biorefinery 10(2),
1170 401-407.
1171 Patwardhan, P.R., Brown, R.C., Shanks, B.H., 2011. Product distribution from the fast pyrolysis of
1172 hemicellulose. ChemSusChem 4(5), 636-643.
1173 Pecha, M.B., Arbelaez, J.I.M., Garcia-Perez, M., Chejne, F., Ciesielski, P.N., 2019. Progress in
1174 understanding the four dominant intra-particle phenomena of lignocellulose pyrolysis: chemical
1175 reactions, heat transfer, mass transfer, and phase change. Green Chemistry 21(11), 2868-2898.
1176 Pemen, A., van Paasen, S., Yan, K., Nair, S., van Heesch, E., Ptasinski, K., Neeft, J., 2002.
1177 Conditioning of biomass derived fuel gas using plasma techniques, Proceedings of the 12th European
1178 Conference on Biomass for Energy, Industry and Climate Protection, Amsterdam, The Netherlands.
1179 pp. 17-21.
1180 Peng, Y., Wu, S., 2010. The structural and thermal characteristics of wheat straw hemicellulose.
1181 Journal of Analytical and Applied Pyrolysis 88(2), 134-139.
1182 Qu, T., Guo, W., Shen, L., Xiao, J., Zhao, K., 2011. Experimental Study of Biomass Pyrolysis Based
1183 on Three Major Components: Hemicellulose, Cellulose, and Lignin. Industrial & Engineering
1184 Chemistry Research 50(18), 10424-10433.
1185 Rabemanolontsoa, H., Saka, S., 2013. Comparative study on chemical composition of various
1186 biomass species. RSC Advances 3(12), 3946-3956.
1187 Radlein, D.S.T.A.G., Grinshpun, A., Piskorz, J., Scott, D.S., 1987. On the presence of anhydro-
1188 oligosaccharides in the sirups from the fast pyrolysis of cellulose. Journal of Analytical and Applied
1189 Pyrolysis 12(1), 39-49.
1190 Rahman, M.A., 2020. Valorizing of weeds algae through the solar assisted pyrolysis: Effects of
1191 dependable parameters on yields and characterization of products. Renewable Energy 147, 937-946.
1192 Raikova, S., Le, C.D., Beacham, T.A., Jenkins, R.W., Allen, M.J., Chuck, C.J., 2017. Towards a
1193 marine biorefinery through the hydrothermal liquefaction of macroalgae native to the United
1194 Kingdom. Biomass and Bioenergy 107, 244-253.
1195 Ranzi, E., Debiagi, P.E.A., Frassoldati, A., 2017. Mathematical Modeling of Fast Biomass Pyrolysis
1196 and Bio-Oil Formation. Note I: Kinetic Mechanism of Biomass Pyrolysis. ACS Sustainable
1197 Chemistry & Engineering 5(4), 2867-2881.
1198 Richter, H., Howard, J.B., 2000. Formation of polycyclic aromatic hydrocarbons and their growth to
1199 soot—a review of chemical reaction pathways. Progress in Energy and Combustion Science 26(4-6),
1200 565-608.

39
1201 Rizzo, A.M., Prussi, M., Bettucci, L., Libelli, I.M., Chiaramonti, D., 2013. Characterization of
1202 microalga Chlorella as a fuel and its thermogravimetric behavior. Applied Energy 102, 24-31.
1203 Rönkkönen, H., Simell, P., Reinikainen, M., Krause, O., Niemelä, M.V., 2010. Catalytic clean-up of
1204 gasification gas with precious metal catalysts–A novel catalytic reformer development. Fuel 89(11),
1205 3272-3277.
1206 Saber, M., Nakhshiniev, B., Yoshikawa, K., 2016. A review of production and upgrading of algal bio-
1207 oil. Renewable and Sustainable Energy Reviews 58, 918-930.
1208 Sadakata, M., Takahashi, K., Saito, M., Sakai, T., 1987. Production of fuel gas and char from wood,
1209 lignin and holocellulose by carbonization. Fuel 66(12), 1667-1671.
1210 Sajdak, M., Muzyka, R., Hrabak, J., Słowik, K., 2015. Use of plastic waste as a fuel in the co-
1211 pyrolysis of biomass: Part III: Optimisation of the co-pyrolysis process. Journal of Analytical and
1212 Applied Pyrolysis 112, 298-305.
1213 Serrano, D., Castelló, D., 2020. Tar prediction in bubbling fluidized bed gasification through artificial
1214 neural networks. Chemical Engineering Journal 402, 126229.
1215 Shafizadeh, F., McGinnis, G.D., Philpot, C.W., 1972. Thermal degradation of xylan and related model
1216 compounds. Carbohydrate Research 25(1), 23-33.
1217 Sharifzadeh, M., Sadeqzadeh, M., Guo, M., Borhani, T.N., Murthy Konda, N.V.S.N., Garcia, M.C.,
1218 Wang, L., Hallett, J., Shah, N., 2019. The multi-scale challenges of biomass fast pyrolysis and bio-oil
1219 upgrading: Review of the state of art and future research directions. Progress in Energy and
1220 Combustion Science 71, 1-80.
1221 Sharma, A., Pareek, V., Zhang, D., 2015. Biomass pyrolysis—A review of modelling, process
1222 parameters and catalytic studies. Renewable and Sustainable Energy Reviews 50, 1081-1096.
1223 Shen, Y., Yoshikawa, K., 2013. Recent progresses in catalytic tar elimination during biomass
1224 gasification or pyrolysis—A review. Renewable and Sustainable Energy Reviews 21, 371-392.
1225 Shin, T.-S., Xue, Z., Do, Y.-W., Jeong, S.-I., Woo, H.-C., Kim, N.-G., 2011. Chemical Properties of
1226 Sea Tangle (Saccharina. japonica) Cultured in the Different Depths of Seawater. Clean Technology
1227 17(4), 395-405.
1228 Shuping, Z., Yulong, W., Mingde, Y., Kaleem, I., Chun, L., Tong, J., 2010. Production and
1229 characterization of bio-oil from hydrothermal liquefaction of microalgae Dunaliella tertiolecta cake.
1230 Energy 35(12), 5406-5411.
1231 Sikarwar, V., Zhao, M., 2016. Design of co-gasification of dried sludge and woody biomass for
1232 synthesis gas production in a fixed bed downdraft gasifier using ASPEN PLUS, Abstracts of Papers
1233 of The Americal Chemical Society. American Chemical Society USA, Philadelphia
1234 Sikarwar, V.S., Hrabovský, M., Van Oost, G., Pohořelý, M., Jeremiáš, M., 2020. Progress in waste
1235 utilization via thermal plasma. Progress in Energy and Combustion Science 81, 100873.

40
1236 Sikarwar, V.S., Ji, G., Zhao, M., Wang, Y., 2017. Equilibrium Modeling of Sorption-Enhanced
1237 Cogasification of Sewage Sludge and Wood for Hydrogen-Rich Gas Production with in Situ Carbon
1238 Dioxide Capture. Industrial & Engineering Chemistry Research 56(20), 5993-6001.
1239 Sikarwar, V.S., Zhao, M., 2017. Biomass Gasification. Encyclopedia of Sustainable Technologies.
1240 Sikarwar, V.S., Zhao, M., Clough, P., Yao, J., Zhong, X., Memon, M.Z., Shah, N., Anthony, E.J.,
1241 Fennell, P.S., 2016. An overview of advances in biomass gasification. Energy & Environmental
1242 Science 9(10), 2939-2977.
1243 Simell, P.A., Hepola, J.O., Krause, A.O.I., 1997. Effects of gasification gas components on tar and
1244 ammonia decomposition over hot gas cleanup catalysts. Fuel 76(12), 1117-1127.
1245 Singh, S., Chakraborty, J.P., Mondal, M.K., 2020. Pyrolysis of torrefied biomass: Optimization of
1246 process parameters using response surface methodology, characterization, and comparison of
1247 properties of pyrolysis oil from raw biomass. Journal of Cleaner Production 272, 122517.
1248 Söyler, N., Goldfarb, J.L., Ceylan, S., Saçan, M.T., 2017. Renewable fuels from pyrolysis of
1249 Dunaliella tertiolecta: An alternative approach to biochemical conversions of microalgae. Energy 120,
1250 907-914.
1251 Su, G., Ong, H.C., Gan, Y.Y., Chen, W.-H., Chong, C.T., Ok, Y.S., 2022. Co-pyrolysis of microalgae
1252 and other biomass wastes for the production of high-quality bio-oil: Progress and prospective.
1253 Bioresource Technology 344, 126096.
1254 Sunphorka, S., Chalermsinsuwan, B., Piumsomboon, P., 2017. Artificial neural network model for the
1255 prediction of kinetic parameters of biomass pyrolysis from its constituents. Fuel 193, 142-158.
1256 Taherzadeh, M.J., Eklund, R., Gustafsson, L., Niklasson, C., Lidén, G., 1997. Characterization and
1257 Fermentation of Dilute-Acid Hydrolyzates from Wood. Industrial & Engineering Chemistry Research
1258 36(11), 4659-4665.
1259 Teixeira, A.R., Gantt, R., Joseph, K.E., Maduskar, S., Paulsen, A.D., Krumm, C., Zhu, C.,
1260 Dauenhauer, P.J., 2016. Spontaneous aerosol ejection: origin of inorganic particles in biomass
1261 pyrolysis. ChemSusChem 9(11), 1322-1328.
1262 Teixeira, A.R., Mooney, K.G., Kruger, J.S., Williams, C.L., Suszynski, W.J., Schmidt, L.D., Schmidt,
1263 D.P., Dauenhauer, P.J., 2011. Aerosol generation by reactive boiling ejection of molten cellulose.
1264 Energy Environmental Science & Technology 4(10), 4306-4321.
1265 Templeton, D.W., Quinn, M., Van Wychen, S., Hyman, D., Laurens, L.M.J.J.o.C.A., 2012. Separation
1266 and quantification of microalgal carbohydrates. 1270, 225-234.
1267 Trendewicz, A., Evans, R., Dutta, A., Sykes, R., Carpenter, D., Braun, R., 2015. Evaluating the effect
1268 of potassium on cellulose pyrolysis reaction kinetics. Biomass and Bioenergy 74, 15-25.
1269 Uddin, M.A., Tsuda, H., Wu, S., Sasaoka, E., 2008. Catalytic decomposition of biomass tars with iron
1270 oxide catalysts. Fuel 87(4-5), 451-459.

41
1271 Uzakov, G.N., Davlonov, H.A., Holikov, K.N., 2018. Study of the Influence of the Source Biomass
1272 Moisture Content on Pyrolysis Parameters. Applied Solar Energy 54(6), 481-484.
1273 Uzun, B.B., Apaydin-Varol, E., Ateş, F., Özbay, N., Pütün, A.E., 2010. Synthetic fuel production
1274 from tea waste: Characterisation of bio-oil and bio-char. Fuel 89(1), 176-184.
1275 Van de Velden, M., Baeyens, J., Brems, A., Janssens, B., Dewil, R., 2010. Fundamentals, kinetics and
1276 endothermicity of the biomass pyrolysis reaction. Renewable energy 35(1), 232-242.
1277 Van Heesch, B.E., Pemen, G., Yan, K., Van Paasen, S.V., Ptasinski, K.J., Huijbrechts, P.A., 2000.
1278 Pulsed corona tar cracker. IEEE Transactions on Plasma Science 28(5), 1571-1575.
1279 Várhegyi, G., Antal, M.J., Jakab, E., Szabó, P., 1997. Kinetic modeling of biomass pyrolysis. Journal
1280 of Analytical and Applied Pyrolysis 42(1), 73-87.
1281 Verma, P., Kumar, M., Mishra, G., Sahoo, D., 2017. Multivariate analysis of fatty acid and
1282 biochemical constitutes of seaweeds to characterize their potential as bioresource for biofuel and fine
1283 chemicals. Bioresource Technology 226, 132-144.
1284 Vuppaladadiyam, A.K., Liu, H., Zhao, M., Soomro, A.F., Memon, M.Z., Dupont, V., 2019a.
1285 Thermogravimetric and kinetic analysis to discern synergy during the co-pyrolysis of microalgae and
1286 swine manure digestate. Biotechnology for Biofuels 12(1), 170.
1287 Vuppaladadiyam, A.K., Memon, M.Z., Ji, G., Raheem, A., Jia, T.Z., Dupont, V., Zhao, M., 2019.
1288 Thermal Characteristics and Kinetic Analysis of Woody Biomass Pyrolysis in the Presence of
1289 Bifunctional Alkali Metal Ceramics. ACS Sustainable Chemistry & Engineering 7(1), 238-248.
1290 Vuppaladadiyam, A.K., Merayo, N., Blanco, A., Hou, J., Dionysiou, D.D., Zhao, M., 2018.
1291 Simulation study on comparison of algal treatment to conventional biological processes for greywater
1292 treatment. Algal Research 35, 106-114.
1293 Vuppaladadiyam, A.K., Zhao, M., Memon, M.Z., Soomro, A.F., 2019b. Microalgae as a renewable
1294 fuel resource: a comparative study on the thermogravimetric and kinetic behavior of four microalgae.
1295 Sustainable Energy & Fuels 3(5), 1283-1296.
1296 Vuppaladadiyam, A.K., Zhao, M., Memon, M.Z., Soomro, A.F., Wei, W., 2019c. Solid waste as a
1297 renewable source of energy: A comparative study on thermal and kinetic behavior of three organic
1298 solid wastes. Energy & Fuels 33(5), 4378-4388.
1299 Wang, S., Cao, B., Abomohra, A.E.-F., Hu, Y., Wang, Q., He, Z., Xu, S., Feng, Y., Bernard, U.B.,
1300 Jiang, X., 2018a. Comparative Study of Combustion Properties of Two Seaweeds in a Batch Fluidized
1301 Bed. Combustion Science and Technology 190(5), 755-769.
1302 Wang, S., Jiang, D., Cao, B., Hu, Y., Yuan, C., Wang, Q., He, Z., Hui, C.-W., Abomohra, A.E.-F.,
1303 Liu, X., Feng, Y., Zhang, B., 2018b. Study on the interaction effect of seaweed bio-coke and rice husk
1304 volatiles during co-pyrolysis. Journal of Analytical and Applied Pyrolysis 132, 111-122.
1305 Wang, S., Jiang, X., Wang, N., Yu, L., Li, Z., He, P., 2007. Research on pyrolysis characteristics of
1306 seaweed. Energy & Fuels 21(6), 3723-3729.
42
1307 Wang, S., Lin, H., Ru, B., Dai, G., Wang, X., Xiao, G., Luo, Z., 2016. Kinetic modeling of biomass
1308 components pyrolysis using a sequential and coupling method. Fuel 185, 763-771.
1309 Wang, S., Ru, B., Lin, H., Luo, Z., 2013. Degradation mechanism of monosaccharides and xylan
1310 under pyrolytic conditions with theoretic modeling on the energy profiles. Bioresource Technology
1311 143, 378-383.
1312 Wang, S., Shang, H., Abomohra, A.E.-F., Wang, Q., 2019. One-step conversion of microalgae to
1313 alcohols and esters through co-pyrolysis with biodiesel-derived glycerol. Energy Conversion and
1314 Management 198, 111792.
1315 Wang, S., Wang, Q., Jiang, X., Han, X., Ji, H., 2013. Compositional analysis of bio-oil derived from
1316 pyrolysis of seaweed. Energy Conversion and Management 68, 273-280.
1317 Wang, X., Bai, S., Jin, Q., Li, S., Li, Y., Li, Y., Tan, H., 2018. Soot formation during biomass
1318 pyrolysis: Effects of temperature, water-leaching, and gas-phase residence time. Journal of Analytical
1319 and Applied Pyrolysis 134, 484-494.
1320 Wang, X., Sheng, L., Yang, X., 2017a. Pyrolysis characteristics and pathways of protein, lipid and
1321 carbohydrate isolated from microalgae Nannochloropsis sp. Bioresource Technology 229, 119-125.
1322 Wang, X., Tang, X., Yang, X., 2017b. Pyrolysis mechanism of microalgae Nannochloropsis sp. based
1323 on model compounds and their interaction. Energy Conversion and Management 140, 203-210.
1324 Ward, J., Rasul, M.G., Bhuiya, M.M.K., 2014. Energy Recovery from Biomass by Fast Pyrolysis.
1325 Procedia Engineering 90, 669-674.
1326 Wei, L., Xu, S., Zhang, L., Zhang, H., Liu, C., Zhu, H., Liu, S., 2006. Characteristics of fast pyrolysis
1327 of biomass in a free fall reactor. Fuel Processing Technology 87(10), 863-871.
1328 Weng, J.-K., Li, X., Bonawitz, N.D., Chapple, C., 2008. Emerging strategies of lignin engineering and
1329 degradation for cellulosic biofuel production. Current Opinion in Biotechnology 19(2), 166-172.
1330 Winkelmann, C., Kuczaj, A.K., Nordlund, M., Geurts, B.J., 2018. Simulation of aerosol formation
1331 due to rapid cooling of multispecies vapors. Journal of Engineering Mathematics 108(1), 171-196.
1332 Xiong, Q., Xu, F., Pan, Y., Yang, Y., Gao, Z., Shu, S., Hong, K., Bertrand, F., Chaouki, J., 2018.
1333 Major trends and roadblocks in CFD-aided process intensification of biomass pyrolysis. Chemical
1334 Engineering and Processing - Process Intensification 127, 206-212.
1335 Xiong, Q., Yang, Y., Xu, F., Pan, Y., Zhang, J., Hong, K., Lorenzini, G., Wang, S., 2017. Overview
1336 of Computational Fluid Dynamics Simulation of Reactor-Scale Biomass Pyrolysis. ACS Sustainable
1337 Chemistry & Engineering 5(4), 2783-2798.
1338 Xu, S., Cao, B., Uzoejinwa, B.B., Odey, E.A., Wang, S., Shang, H., Li, C., Hu, Y., Wang, Q.,
1339 Nwakaire, J.N., 2020. Synergistic effects of catalytic co-pyrolysis of macroalgae with waste plastics.
1340 Process Safety and Environmental Protection 137, 34-48.

43
1341 Yadavalli, R., Rao, C., S, R.R., Potumarthi, R., 2014. Dairy effluent treatment and lipids production
1342 by Chlorella pyrenoidosa and Euglena gracilis: Study on open and closed systems. Asia-Pacific
1343 Journal of Chemical Engineering 9.
1344 Yang, C., Li, R., Qiu, Q., Yang, H., Zhang, Y., Yang, B., Wu, J., Li, B., Wang, W., Ding, Y., Zhang,
1345 B., 2020. Pyrolytic behaviors of Scenedesmus obliquus over potassium fluoride on alumina. Fuel 263,
1346 116724.
1347 Yang, C., Li, R., Zhang, B., Qiu, Q., Wang, B., Yang, H., Ding, Y., Wang, C., 2019. Pyrolysis of
1348 microalgae: A critical review. Fuel Processing Technology 186, 53-72.
1349 Yang, H., Li, S., Liu, B., Chen, Y., Xiao, J., Dong, Z., Gong, M., Chen, H., 2020. Hemicellulose
1350 pyrolysis mechanism based on functional group evolutions by two-dimensional perturbation
1351 correlation infrared spectroscopy. Fuel 267, 117302.
1352 Yang, H., Yan, R., Chen, H., Lee, D.H., Zheng, C., 2007. Characteristics of hemicellulose, cellulose
1353 and lignin pyrolysis. Fuel 86(12-13), 1781-1788.
1354 Yang, M., Luo, B., Shao, J., Zeng, K., Zhang, X., Yang, H., Chen, H., 2018. The influence of CO2 on
1355 biomass fast pyrolysis at medium temperatures. Journal of Renewable and Sustainable Energy 10(1),
1356 013108.
1357 Yazdani, E., Hashemabadi, S.H., Taghizadeh, A., 2019. Study of waste tire pyrolysis in a rotary kiln
1358 reactor in a wide range of pyrolysis temperature. Waste Management 85, 195-201.
1359 Yoon, K., Lee, S.S., Ok, Y.S., Kwon, E.E., Song, H., 2019. Enhancement of syngas for H2 production
1360 via catalytic pyrolysis of orange peel using CO2 and bauxite residue. Applied Energy 254, 113803.
1361 Yuan, C., Wang, S., Qian, L., Barati, B., Gong, X., Abomohra, A.E.-F., Wang, X., Esakkimuthu, S.,
1362 Hu, Y., Liu, L., 2019. Effect of cosolvent and addition of catalyst (HZSM-5) on hydrothermal
1363 liquefaction of macroalgae. 43(14), 8841-8851.
1364 Zeng, K., Gauthier, D., Li, R., Flamant, G., 2015a. Solar pyrolysis of beech wood: Effects of pyrolysis
1365 parameters on the product distribution and gas product composition. Energy 93, 1648-1657.
1366 Zeng, K., Minh, D.P., Gauthier, D., Weiss-Hortala, E., Nzihou, A., Flamant, G., 2015b. The effect of
1367 temperature and heating rate on char properties obtained from solar pyrolysis of beech wood.
1368 Bioresource Technology 182, 114-119.
1369 Zeng, X., Ueki, Y., Yoshiie, R., Naruse, I., Wang, F., Han, Z., Xu, G., 2020. Recent progress in tar
1370 removal by char and the applications: A comprehensive analysis. Carbon Resources Conversion 3, 1-
1371 18.
1372 Zhang, N., Duan, H., Miller, T.R., Tam, V.W.Y., Liu, G., Zuo, J., 2020. Mitigation of carbon dioxide
1373 by accelerated sequestration in concrete debris. Renewable and Sustainable Energy Reviews 117.
1374 Zhao, B., Xu, X., Li, H., Chen, X., Zeng, F., 2018. Kinetics evaluation and thermal decomposition
1375 characteristics of co-pyrolysis of municipal sewage sludge and hazelnut shell. Bioresource technology
1376 247, 21-29.
44
1377 Zheng, Y., Tao, L., Yang, X., Huang, Y., Liu, C., Zheng, Z., 2019. Comparative study on pyrolysis
1378 and catalytic pyrolysis upgrading of biomass model compounds: Thermochemical behaviors, kinetics,
1379 and aromatic hydrocarbon formation. Journal of the Energy Institute 92(5), 1348-1363.
1380 Zhong, H., Xiong, Q., Zhu, Y., Liang, S., Zhang, J., Niu, B., Zhang, X., 2019. CFD modeling of the
1381 effects of particle shrinkage and intra-particle heat conduction on biomass fast pyrolysis. Renewable
1382 Energy 141, 236-245.
1383 Zhong, L., Zhang, J., Ding, Y., 2020. Energy Utilization of Algae Biomass Waste Enteromorpha
1384 Resulting in Green Tide in China: Pyrolysis Kinetic Parameters Estimation Based on Shuffled
1385 Complex Evolution. Sustainability 12(5), 2086.
1386 Zhou, S., Pecha, B., van Kuppevelt, M., McDonald, A.G., Garcia-Perez, M., 2014. Slow and fast
1387 pyrolysis of Douglas-fir lignin: Importance of liquid-intermediate formation on the distribution of
1388 products. Biomass and Bioenergy 66, 398-409.

1389

45
Highlights
• Biomass pyrolysis mechanisms concerning feedstock type are critically discussed
• Pyrolysis mechanisms play a major role in deciding the quality of end products
• Recent advancements in modeling ease the understanding of pyrolysis mechanisms
• Aerosol formation during pyrolysis severely impacts the quality of bio-oil

You might also like