You are on page 1of 14

Bioresource Technology 344 (2022) 126096

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Review

Co-pyrolysis of microalgae and other biomass wastes for the production of


high-quality bio-oil: Progress and prospective
Guangcan Su a, Hwai Chyuan Ong b, c, *, Yong Yang Gan a, Wei-Hsin Chen d, Cheng Tung Chong e,
Yong Sik Ok f
a
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia
b
Centre for Green Technology, Faculty of Engineering and IT, University of Technology Sydney, NSW 2007, Australia
c
Future Technology Research Center, National Yunlin University of Science and Technology, 123 University Road, Section 3, Douliou, Yunlin 64002, Taiwan
d
Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 701, Taiwan
e
China-UK Low Carbon College, Shanghai Jiao Tong University, Lingang, Shanghai 201306, PR China
f
Korea Biochar Research Center, APRU Sustainable Waste Management Program & Division of Environmental Science and Ecological Engineering, Korea University,
Seoul 02841, Republic of Korea

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Synergistic effects of microalgae co-


pyrolysis has been critically analyzed.
• Co-pyrolysis enhances the overall py­
rolysis performance and bio-oil
properties.
• CaO and Cu/HZSM-5 promote the
quality of bio-oil effectively.
• Co-pyrolysis of microalgae and other
biomass wastes is economic and
feasible.

A R T I C L E I N F O A B S T R A C T

Keywords: Microalgae are the most prospective raw materials for the production of biofuels, pyrolysis is an effective method
Microalgae to convert biomass into bioenergy. However, biofuels derived from the pyrolysis of microalgae exhibit poor fuel
Biofuels properties due to high content of moisture and protein. Co-pyrolysis is a simple and efficient method to produce
Co-pyrolysis
high-quality bio-oil from two or more materials. Tires, plastics, and bamboo waste are the optimal co-feedstocks
Catalytic effect
based on the improvement of yield and quality of bio-oil. Moreover, adding catalysts, especially CaO and Cu/
Environmental sustainability
HZSM-5, can enhance the quality of bio-oil by increasing aromatics content and decreasing oxygenated and
nitrogenous compounds. Consequently, this paper provides a critical review of the production of bio-oil from co-
pyrolysis of microalgae with other biomass wastes. Meanwhile, the underlying mechanism of synergistic effects
and the catalytic effect on co-pyrolysis are discussed. Finally, the economic viability and prospects of microalgae
co-pyrolysis are summarized.

* Corresponding author.
E-mail addresses: ong1983@yahoo.com, HwaiChyuan.Ong@uts.edu.au (H.C. Ong).

https://doi.org/10.1016/j.biortech.2021.126096
Received 16 August 2021; Received in revised form 2 October 2021; Accepted 4 October 2021
Available online 6 October 2021
0960-8524/© 2021 Elsevier Ltd. All rights reserved.
G. Su et al. Bioresource Technology 344 (2022) 126096

1. Introduction emission hamper its development. Dry or wet torrefaction can transform
microalgae biomass into solid product, but it is used as a pretreatment
With the rapid development of industrialization and urbanization, method to reduce the moisture content and improve grinding and hy­
the energy demand is growing dramatically, while the finite conven­ drophobicity of biomass for pyrolysis (He et al., 2018). The combustion
tional fossil fuels cannot meet the requirement (Liu et al., 2020a). Based of microalgae can produce heat or electricity, but low thermal efficiency
on the Annual Energy Outlook 2020 from the United States Energy In­ and the release of noxious gas and ash hinder its further development.
formation Agency, crude oil and coal reserves are estimated to be Hydrothermal liquefaction can directly convert wet microalgae at a
depleted in 2042 and 2112, respectively, as global energy demand will moderate temperature (300–400 ◦ C) and high pressure environment
rise by 56% in 2040 (EIA, 2020). Furthermore, the reckless consumption (10–25 MPa) into bio-oil, biochar, and biogas without any pretreatment
of natural resources and the uninterrupted expenditure of fossil fuels process. Nevertheless, huge capital costs, low security, and unobserv­
have led to a series of grave problems including air pollution, global able reaction process impede its promotion (Chiaramonti et al., 2017).
warming, acid rain, ozone layer loss, and desertification, which have Pyrolysis is regarded as the most popular thermochemical conver­
constituted an enormous threat to the existence and development of sion technology, which thermally degrades the organic materials by the
mankind. It is urgent and essential to search for clean, alternative, and cracking of chemical bonds in the absence of oxygen. Pyrolysis is cate­
sustainable energy to reduce the degree of reliance on fossil fuels (Uzar, gorized as slow, fast, and flash pyrolysis based on various operating
2020). conditions including temperature, heating rate, and residence time,
Renewable and sustainable energy sources such as solar, wind, nu­ whereas the product distribution can be highly controllable by tailoring
clear, geothermal, and hydro have been successfully utilized to decrease those pyrolysis parameters (Ong et al., 2019). Moreover, it is a cheap
the consumption and dependency on fossil fuels (Benedek et al., 2018). and effective route to produce biofuels and valuable chemicals because
However, those energy sources are intermittent and non-dispatchable, of the superiorities of abundant feedstocks, unstrict conditions, high
their output power is hard to satisfy the fluctuating electricity needs conversion efficiency, and eco-friendliness, thus it is on the edge of
of society at any time owing to the change of power generation. Bio­ commercialization right now. However, owing to the high oxygen and
energy originated from biomass has overcome the above problem, the nitrogen contents in bio-oil, some barriers have impeded the process,
primary products including biochar and biofuels, can be used as vital such as low caloric value, high acidity, low stability, nitrogen pollution,
and stable backup energy for storage and transportation (Chen et al., and corrosiveness. To eliminate the obstacles and enhance the quality of
2021b). Moreover, biomass is the most promising energy source, which bio-oil, many researchers have explored in-depth and focused on co-
has drawn attention globally because of its low cost, sustainability, pyrolysis, proving that co-pyrolysis is an efficient way to solve the
availability, and environmental friendliness (Ong et al., 2020). Various above problems (Su et al., 2021).
kinds of materials have been employed as feedstocks for the production Co-pyrolysis involves two or more different materials, the interaction
of biofuels, such as corn, wood, wheat straw, sewage sludge, plastics, between various radicals possesses a promotional effect on the forma­
tires, kitchen waste, and microalgae (Lu et al., 2020). Currently, biofuels tion of homogenous bio-oil and leads to synergistic effects, which are the
are classified into four generations according to different biomass crux to improve the properties of biofuels. During the co-pyrolysis
sources. First-generation biofuels are mainly manufactured from animal process, tires, plastics, and bamboo waste have exhibited the best syn­
fats or oil-based plants, but the application causes many serious conse­ ergistic effects on the enhancement of yield and higher heating value
quences like water crises, ecosystem damage, and food versus fuel (HHV) of bio-oil with microalgae (Dai et al., 2018b). Meanwhile, due to
debate (Stamenković et al., 2020). Second-generation biofuels have the significant advantages of simplicity and effectiveness, co-pyrolysis is
been extensively implemented by exploiting agricultural, industrial, also treated as an ideal waste management technology. The utilization of
municipal, and household waste as feedstocks to overcome the problems waste materials including rural solid waste, kitchen waste, wood
raised by the first generation biofuels. However, there are still some sawdust, peanut shell, and sewage sludge as co-feedstocks with micro­
commercial restrictions, especially the high cost of transportation, pre­ algae has also achieved remarkable results (Bong et al., 2020; Chang
treatment, and conversion (Lee et al., 2015). Third-generation biofuels et al., 2018; Chen et al., 2018a). Furthermore, plenty of studies have
are derived from microalgae, which exhibit numerous advantages indicated that catalysts can further facilitate the selectivity of the py­
including high photosynthetic efficiencies, good CO2 capture ability, rolysis process and the properties of bio-oil by limiting product distri­
high oil content, exponential growth rates, and non-selective growth bution, lowering the activation energy, reducing acidity, and cutting
condition. Thus, microalgae have been suggested as the most promising down the oxygenated and nitrogenous compounds (Xiang et al., 2018).
and sustainable raw materials for the production of biofuels, pharma­ As the enormous advantages of microalgae and catalytic co-pyrolysis
ceuticals, chemicals, and other high-value products, but they are still process, catalytic co-pyrolysis of microalgae and other biomass wastes
constrained by low productivity and high cost of cultivation and harvest will have a crucial influence on the energy market and play a key role in
(Gan et al., 2018). Additionally, fourth-generation biofuels employ the future. Numerous literature works on the conventional pyrolysis, co-
genetically modified microalgae to produce biofuels. Genetic modifi­ pyrolysis, hydrothermal liquefaction, and catalytic pyrolysis of micro­
cation is utilized to increase the output of microalgae by raising the algae have been published (Chen et al., 2015; Enamala et al., 2018;
photosynthetic efficiency, decreasing the photoinhibition, and Fakayode et al., 2020). Nevertheless, a review of the non-catalytic and
improving the light penetration. While it is an emerging technology, catalytic co-pyrolysis of microalgae and other biomass wastes is limited
which is in its infancy and requires more in-depth research (Abdullah in the literature, even with the above advantages. Consequently, this
et al., 2019). paper reveals a critical and comprehensive review of the co-pyrolysis of
Simultaneously, a proper treatment process is required to utilize microalgae with other biomass wastes. Furthermore, the mechanism of
microalgae as an energy source. There are two main categories of synergistic effects between various materials and the catalytic effect on
biomass conversion methods including thermochemical conversion co-pyrolysis are provided. Finally, the economic viability, challenges,
(pyrolysis, torrefaction, combustion, gasification, and liquefaction) and and prospects of co-pyrolysis are fully discussed.
biochemical conversion (fermentation and digestion) (Lee et al., 2020b).
Presently, many conversion techniques, such as gasification, torre­ 2. Microalgae and other biomass wastes characterization
faction, combustion, and hydrothermal liquefaction, have been exten­
sively explored. Gasification is a versatile thermochemical technology, As the oldest living unicellular microorganisms (1–400 μm) on earth,
which utilizes air, oxygen, steam, or supercritical water to convert microalgae have been available for about 3 billion years with approxi­
microalgae biomass into syngas (H2, CO, CO2, and CH4) at a temperature mately 20,000 species (Levasseur et al., 2020). Moreover, microalgae
range of 700 to 1000 ◦ C. However, high energy input and toxic gases have relatively low requirements for growth environment, which can be

2
G. Su et al. Bioresource Technology 344 (2022) 126096

Table 1
Elemental and proximate analyses of microalgae and other biomass wastes.
Biomass Elemental Analysis (dry basis, wt%) Proximate Analysis (wt%) HHV(MJ/kg) Reference
C H O N S VM MO FC Ash

Microalgae
Chlorella vulgaris 46.82 8.63 32.03 11.05 0.29 77.23 1.98 12.35 8.44 22.82 (Chen et al., 2018a)
Nannochloropsis sp. 53.39 7.75 21.27 6.97 0.61 79.61 4.01 10.38 6.00 24.01 (Chen et al., 2017a)
Nannochloropsis sp. 50.61 7.31 28.46 6.68 0.64 79.21 4.23 10.26 6.30 21.51 (Tang et al., 2020b)
Isochrysis sp. 49.26 7.50 31.74 6.24 0.96 79.79 1.69 11.63 6.89 23.52 (Wang et al., 2016)
Spirulina platensis 47.27 7.05 21.53 10.99 0.62 82.03 7.48 5.43 5.06 21.21 (Chen et al., 2018c)
Spirulina platensis 47.58 6.33 28.15 11.26 0.51 81.13 7.62 5.08 6.17 18.68 (Tang et al., 2020b)
Scenedesmus sp. 45.60 6.50 36.30 11.10 0.48 75.20 12.10 7.10 5.60 23.40 (Nyoni et al., 2020)
Dunaliella salina 48.10 7.10 23.30 9.40 0.90 76.30 4.00 12.50 7.20 21.20 (Gong et al., 2014)
Dunaliella salina 45.75 6.91 43.08 4.11 0.15 81.80 – 17.98 0.22 – (Chen et al., 2021a)
Mougeotia 41.51 5.59 27.03 5.40 0.51 64.49 5.14 10.41 19.96 16.63 (Sharara et al., 2014)
Arthrospira platensis 46.75 6.97 35.11 10.62 0.55 71.50 7.60 15.93 4.97 20.75 (Anand et al., 2016)
Dunaliella tertiolecta 43.31 5.96 29.90 4.33 – 64.70 1.60 17.20 16.50 17.81 (Wu et al., 2015)
Isochrysis galbana 48.35 7.24 36.82 7.60 – 64.02 2.27 10.12 23.56 16.52 (Azizi et al., 2020)
Botryococcus braunii 51.50 8.90 31.90 7.50 – 76.10 3.50 17.90 2.40 21.50 (Ali et al., 2018)
Cyanobacteria sp. 33.56 5.16 25.89 5.85 0.62 61.78 – 9.32 28.92 14.48 (Huang et al., 2016)
Chlorella sp. 49.00 7.14 28.97 8.59 0.77 79.17 – 15.41 5.42 – (Lv et al., 2021)
Chroococcus sp. 24.28 3.98 25.35 4.17 1.08 49.23 6.77 4.20 39.80 – (Varsha et al., 2021)
Lignocellulosic biomass
Bamboo 41.97 5.89 43.18 0.27 0.15 80.57 6.22 10.89 2.32 16.86 (Chen et al., 2017a)
Wood sawdust 47.10 6.63 46.13 0.15 0.00 79.76 5.32 14.21 0.72 18.42 (Chen et al., 2019b)
Peanut crust 43.80 5.70 34.00 1.80 0.40 72.40 5.80 13.20 8.50 18.30 (Sattar et al., 2020)
Palm kernel shell 50.70 6.00 42.80 0.40 0.10 75.20 0.00 22.70 2.10 – (Chang et al., 2018)
Walnut shell 61.09 6.42 29.81 0.16 0.16 74.53 1.06 22.05 2.36 – (Zhu et al., 2018)
Cornstalk 42.89 5.94 43.11 0.91 – 72.74 3.46 16.65 7.15 – (Li et al., 2017)
Non-lignocellulosic biomass
Rural solid waste 51.03 7.43 27.93 1.31 0.14 75.26 0.00 10.64 12.16 22.39 (Tang et al., 2020a)
Kitchen waste 45.15 8.57 36.31 1.53 0.00 80.27 2.47 16.07 1.19 18.74 (Chen et al., 2018a)
Municipal solid waste 55.67 7.57 33.99 2.77 0.43 75.73 7.13 6.38 16.75 – (Varsha et al., 2021)
Low-quality coal 79.31 4.72 13.38 1.03 1.30 30.56 4.18 49.88 15.38 – (Wu et al., 2018a)
Semi-anthracite coal 63.15 4.74 8.84 0.78 0.92 33.12 0.00 45.31 21.57 26.32 (Chen et al., 2012)
Colombian coal 79.50 5.00 13.20 1.60 0.00 43.20 0.00 56.10 0.70 32.20 (Fermoso et al., 2018)
Coal 63.60 3.64 30.70 1.57 0.49 20.90 2.80 60.10 16.20 27.80 (Nyoni et al., 2020)
LDPE 84.32 15.47 0.00 0.00 0.21 99.99 0.00 0.00 0.00 47.29 (Tang et al., 2020b)
PVC 38.09 5.22 – – – 93.77 0.00 6.06 0.17 20.07 (Dai et al., 2018b)
PP 84.31 15.30 0.00 0.31 0.08 99.60 0.40 0.00 0.00 47.46 (Azizi et al., 2018b)
PS 92.07 7.89 0.00 0.01 0.03 99.57 – 0.39 0.04 – (Chen et al., 2021a)
PET 61.80 4.55 33.55 0.02 0.08 84.80 0.00 14.80 0.40 – (Chen et al., 2021a)
Waste rubber tires 78.70 7.47 1.63 1.40 1.90 72.00 0.00 19.00 9.00 36.09 (Duan et al., 2015)
Tire 78.95 6.73 1.62 0.72 1.78 60.33 0.00 29.47 10.20 – (Fang et al., 2018)
Waste cooking oil 78.31 11.46 10.01 0.00 0.08 97.57 – 0.39 0.14 39.93 (Wu et al., 2020)
Waste lubricant oil 83.90 14.20 0.44 0.70 0.76 99.09 – – 0.91 43.10 (Uçar et al., 2005)
Oil shale 37.34 4.90 51.99 1.47 4.30 38.52 3.63 5.38 52.47 – (Hu et al., 2016)
Sewage sludge 34.77 4.27 24.69 2.59 0.00 60.22 2.44 2.49 34.85 13.04 (Wang et al., 2016)
Textile dyeing sludge 24.75 3.80 3.93 3.26 10.44 41.33 – 4.85 53.82 – (Peng et al., 2015)

VM: Volatile matter; MO: Moisture; FC: Fixed carbon; HHV: Higher heating value (MJ/kg); LDPE: Low-density polyethylene; PP: Polypropylene; PS: Polystyrene; PET:
Polyethylene terephthalate.

cultivated in freshwater, brackish water, marine area, and even land. In addition, as the low-cost and most abundant carbon source,
With the help of chloroplasts, microalgae possess the ability to produce lignocellulosic biomass is deemed to be an economically feasible,
rich organic matters, fix carbon, and release oxygen with high photo­ environmental-friendly, and carbon–neutral raw material for the pro­
synthetic efficiency, which are treated as a promising carbon sink (Saj­ duction of biofuels (Dai et al., 2020). Currently, there are about 5 billion
jadi et al., 2018). Microalgae are mainly composed of lipids, proteins, metric tons of crop residues generated across the world annually. In
and carbohydrates, with other minor components including vitamins, contrast to microalgae, the chief components of lignocellulosic biomass
fatty acids, pigments, and antioxidants. The content of major compo­ are cellulose (40–60%), hemicellulose (15–30%), and lignin (10–25%),
nents is primarily related to the species, cultivation methods, and with little extractives and ash, and the composition mainly depends on
environmental conditions (Nagappan et al., 2019). The proteins content the type of source (softwood, hardwood, or herbaceous plant) (Sankaran
of microalgae is in the interval of 14.00 to 65.20 wt%, which is widely et al., 2020). Meanwhile, more attention has been paid to non-
used in the field of pharmacy vaccine, nutrition, and adhesives (Mene­ lignocellulosic biomass such as municipal solid waste, plastics, tires,
gazzo & Fonseca, 2019). Lipids are the most valuable components with coal, and sewage sludge. The global municipal solid waste has increased
the content of 1.10–51.00 wt%, which are suitable for the production of to 2.01 billion tons in 2016, which is estimated to rise to about 3.4
high-value oil-based products (Katiyar & Arora, 2020). Moreover, the billion tons by 2050 (Istrate et al., 2020). In addition, over 300 million
content of carbohydrates is between 3.28 and 30.21 wt%, which are the tons of plastics are produced annually, resulting in plenty of plastics
feedstock for valuable chemicals (Klein et al., 2018). The microalgae waste (Kasar et al., 2020). Furthermore, non-biodegradable waste tires
production industry still faces several challenges like enormous capital are another global environmental problem, more than 1.5 billion tires
investments, high running costs, and huge energy consumption, espe­ are scrapped annually (Mohajerani et al., 2020). The disposal of these
cially in the cultivation and harvesting process. In 2019, the cost of wastes is a huge global challenge, which possesses a serious threat to the
biocrude based on microalgae is approximately $7.50 per gallon, which environment and human survival without proper treatment. Co-
is much higher than conventional fuels (Yang et al., 2019). pyrolysis is an effective and simple solution, it not only can reduce the

3
G. Su et al. Bioresource Technology 344 (2022) 126096

amount of waste, but also generates valuable products. anaerobic condition, a range of chemical reactions, such as decarbox­
Table 1 presents the elemental and proximate analyses of various ylation, decarbonylation, dehydration, and dehydrogenation, occur to
kinds of materials. Most microalgae species possess high nitrogen and form carbonyl, carboxyl, hydroxyl, and other functional groups. Biomass
oxygen content. The content of nitrogen in microalgae is much higher mixtures are then converted into biochar and biogas, while some volatile
than other materials owing to the existence of protein, Spirulina platensis components are condensed into liquid products (Tripathi et al., 2016).
possesses the highest nitrogen content of 11.26 wt%. Meanwhile, Pyrolysis is highly affected by various operating parameters, such as
microalgae species are medium in volatile matter, moisture, fixed car­ temperature, reaction time, heating rate, types of inert gas, sweeping
bon, ash, and HHV. Moreover, the oxygen content of lignocellulosic rate, and particulate size. While co-pyrolysis possesses an extra impor­
biomass is also high, while the carbon and hydrogen content are rela­ tant interfering factor, that is blending ratio of different raw materials.
tively low and lead to low HHV, the nitrogen and sulfur content are close Furthermore, those conditions require to be adjusted according to the
to zero. For non-lignocellulosic biomass, polypropylene (PP), coal, and characteristic of feedstocks and the optimization of the target product.
tires possess high carbon content of 84.31, 79.31, and 78.70 wt%, Condensation is another vital step to isolate bio-oil. In a typical
respectively. Additionally, PP possesses the highest hydrogen content, workflow of the fixed-bed reactor, pyrolysis volatiles are condensed into
volatile matter, HHV, and the lowest ash content. Sewage sludge has the bio-oil by a water-cooled condenser, solid residues are left in the reactor,
highest ash content and the lowest HHV, while coal has the maximum while the non-condensable gases are collected in a gas bag or released
fixed carbon content and the minimum volatile matter content. into the atmosphere after filtration. Three stages condensation system is
widely employed during the pyrolysis experiment (He et al., 2019).
3. Co-pyrolysis of microalgae and other biomass wastes
3.2. Synergistic effects on the co-pyrolysis process
Co-pyrolysis possess the superiorities of simplicity and effectiveness,
while microalgae exhibit the strengths of environmental friendliness, Synergistic effects are the crux to modify the existing pyrolysis
high photosynthetic efficiencies, high oil content, and exponential process and ameliorate the properties of bio-oil. The sufficient contact of
growth rates. Therefore, the combination of co-pyrolysis and microalgae volatiles produced by each material plays a crucial role in synergistic
will have a huge impact on the biomass industry, while numerous re­ effects during the co-pyrolysis process. A large number of research
searchers at home and abroad have devoted themselves to related groups have delved into the phenomenon, especially on the mechanism
studies and yielded fruitful results. of synergistic effects and the improvement of target product during
microalgae co-pyrolysis (Sun et al., 2019a).
3.1. Co-pyrolysis process
3.2.1. Mechanism of synergistic effects
Co-pyrolysis entails the thermal degradation of two or more mate­ The underlying mechanism of synergistic effects between microalgae
rials. It is analogous to conventional pyrolysis which is carried out in a and plastics has been explored and summarized by Tang et al. (2020b)
closed reactor system by utilizing moderate operating temperature and and Qi et al. (2018). With the aid of pyrolysis–gas chromatography-mass
oxygen-free environment. The transition from pyrolysis to co-pyrolysis spectrometry, Tang et al. (2020b) detected that a strong interaction
does not need any modification of existing equipment but positively between microalgae and low-density polyethylene (LDPE) promoted the
influences the pyrolysis process and end-products (Lee et al., 2020a). formation of esters and long-chain alcohols, whereas hindered the pro­
The whole process of co-pyrolysis is divided into three major steps duction of oxygenated and nitrogenous compounds in the bio-oil. The
including pretreatment, co-pyrolysis, and condensation. Pretreatment is free radicals from LDPE enhanced the dehydroxylation and C− C bond
a pivotal step in biomass conversion due to the complex nature and cleavage of long-chain carboxylic acids from the hydrolysis of proteins
structure of biomass materials, which can improve the usability of and lipids. Hydration, decarboxylation, and dehydration reforming re­
components for an extensive range of applications (Soltanian et al., actions made carbohydrates generate formic acid, acetic acids, OH
2020). Pretreatment is classified as physical (milling, freezing, and radicals, and release H2O, CO2, hinder the formation of furans, phenol,
extrusion pretreatments), chemical (acid, alkaline, organic solvent, and long-chain fatty acids. Meanwhile, OH radicals reacted with formic
oxidative, and ionic liquid pretreatments), physical–chemical (steam acid and acetic acids to form esters, or further combined with the long-
explosion, liquid hot water, wet oxidation, CO2 explosion, and ultra­ chain CH radicals from the cleavage of LDPE to generate long-chain
sound pretreatments), biological (microbial consortium, bacterial, alcohols. Moreover, Qi et al. (2018) observed a sharp rise of aromatic
fungal, and enzymatic pretreatments), thermal (torrefaction, drying, hydrocarbons content in the bio-oil due to violent synergistic effects
and microwave irradiation) and combined pretreatments (Huang et al., between Nannochloropsis sp. and PP with the help of HZSM-5. There are
2020). Drying, torrefaction, milling, and sieving are the most common three major routes of the mechanism for the production of aromatic
pretreatment methods to obtain small and homogenous samples. High hydrocarbons. Firstly, olefins from the pyrolysis of PP reacted with fu­
moisture content in biomass can affect the quality of bio-oil negatively. rans from the cracking of microalgae carbohydrates to generate aro­
Thus, drying is the most common technique to reduce the moisture matic hydrocarbons by the Diels-Alder and dehydration reactions, then
content of biomass (Carvalho et al., 2020). Meanwhile, torrefaction can further formed compounds with six-membered ring structures on the
also alter the structure and composition of biomass and improve its surface of HZSM-5. Secondly, small organic moieties were deoxygenated
grinding, hydrophobicity, and pyrolysis performance (He et al., 2018). and decomposed into olefins, while triglycerides were cracked into
Additionally, small particles are conducive to ameliorate the reaction heavy oxygenated hydrocarbons, then underwent decarboxylation and
efficiency and promote the production of biofuels on the basis of surface decarbonylation reactions to form olefins. Moreover, with the help of
chemistry. Therefore, dry biomass particles with the moisture content of HZSM-5, the volatiles originated from the pyrolysis of microalgae and
less than 10 wt% need to be smaller than 1 mm to achieve the optimum the olefins generated from the cracking of PP entered the hydrocarbon
effect of fast pyrolysis, while flash pyrolysis requires even smaller pool and were converted into aromatic hydrocarbons through the hy­
biomass particles (less than 0.2 mm) (Kumar et al., 2020). In general, drocarbon pool mechanism. Thirdly, olefins generated from the pyrol­
pretreatment is the most expensive process because of large energy ysis of PP were transformed into aromatic hydrocarbons through HZSM-
consumption, different kinds of pretreatment should be selected based 5 via aromatization, cyclization, and oligomerization reactions.
on the properties of biomass, capital cost, production condition, and According to Table 1, microalgae have a very high content of ni­
environmental impacts (del Río et al., 2020). trogen, which will cause high nitrogen content in the bio-oil and sub­
During the co-pyrolysis process, biomass mixtures are heated to a sequent nitrogen pollution. Thus, it is essential to inhibit the formation
temperature that exceeds their thermal stability threshold under of nitrogenous compounds (Liu et al., 2016). Meanwhile, proteins are

4
G. Su et al. Bioresource Technology 344 (2022) 126096

Table 2
Synergistic effects of the co-pyrolysis of microalgae with other biomass wastes.
Microalgae Co- Conditions Reactor Relevant results and observation Reference
feedstock

Nannochloropsis sp. Bamboo 600 ◦ C, 30 min, Ar Fixed-bed The increase of microalgae ratio enhances the content of long-chain fatty acids, (Chen et al.,
reactor nitrogenous compounds, decreases the content of phenol, acetic acids, and 2017a)
oxygenated compounds dramatically in the bio-oil.
Nannochloropsis sp. Palm kernel 600 ◦ C, 15 min, N2 Fixed-bed The increase of microalgae ratio decreases the content of phenolic compounds, (Chang et al.,
shell reactor while the content of aromatic hydrocarbons increases first and then decreases. 2018)
Chlorella vulgaris Waste 900 ◦ C, 60 min, N2 Fixed-bed Co-pyrolysis with 50 wt% waste glycerol increases the content of alcohol to (Wang et al.,
glycerol reactor 72.30 wt%, decreases the content of nitrogenous compounds and phenols. 2019a)
Chlorella vulgaris Kitchen 500–700 ◦ C, 25 Fixed-bed The increment of microalgae ratio enhances the hydrocarbons content and (Chen et al.,
waste min, CO2 reactor activation energy, decreases the carboxylic acids, causes a higher bio-oil HHV. 2018a)
Chlorella vulgaris PVC 800 W/1 kW, 40 Microwave The growth of PVC ratio boosts the aromatic hydrocarbons content and HHV, (Dai et al.,
min, N2 reactor decreases oxygenated and nitrogenous compounds in the bio-oil. 30 wt% PVC 2018b)
is the optimum ratio.
Nannochloropsis sp. LDPE 600 ◦ C, 30 min, Ar Fixed-bed The addition of LDPE with high EHI reduces the content of oxygenated and (Tang et al.,
reactor nitrogenous compounds. The aliphatic hydrocarbons content in the oil rises to 2019)
77.40 wt% with 25 wt% LDPE.
Isochrysis sp. Sewage 5, 10, 15, 20 and Thermal Significant synergistic effects are observed at 550 to 700 ◦ C. The addition of (Wang et al.,
sludge 25 ◦ C/min, N2 analyzer microalgae decreases the content of hydrocarbons and nitrogenous 2016)
compounds, increases the content of aldehyde and ketones in the bio-oil.
Dunaliella salina PP, PS, PET, 800 ◦ C, 20 ◦ C/min, Thermal The coating effect of the molten plastics enhances the pyrolysis of microalgae; (Chen et al.,
PVC N2 analyzer the hydrogenation reaction between microalgae and PP, PS, and PET reduces 2021a)
the solid residue; the activation energies of PP, PS, PET, and PVC are decreased
largely after adding microalgae.
Chlorella sp. PP 550 ◦ C, 20 s, Ar Fixed-bed The addition of PP reduces the activation energy remarkably; The increase in (Lv et al.,
reactor the PP ratio in the mixture raises the hydrocarbons content and decreases the 2021)
oxygenated compounds largely.
Chlorella, Low-rank 25–850 ◦ C, Thermal The presence of microalgae enhances the release of volatiles from coal; (Wu et al.,
Nannochloropsis coal 10–40 ◦ C/min, N2 analyzer Different species of microalgae influences the decomposition of coal distinctly; 2018b)
sp. synergistic effects cut down the activation energy.
Chlorella pyrenoidosa Waste 290–370 ◦ C, Mini-batch 50% microalgae and 50% waste tire induce the most violent synergistic effects, (Duan et al.,
rubber tire 10–120 min, He reactor which promote denitrogenation and deoxygenation, thereby improving the 2015)
quality of bio-oil.

EHI: Effective hydrogen index.

the primary component of microalgae and the chief nitrogen source, microalgae underwent C− O cleavage through acyl or alkoxy during
while amino acids are the basic structural units of protein. Therefore, the pyrolysis. Therefore, pyrolysis can enhance the cleavage of the alkoxy
mechanism of synergy between microalgae and lignocellulosic biomass bond and generate a great number of long-chain fatty acids.
can be revealed through the co-pyrolysis of amino acids and lignocel­ Other researchers reported that the synergistic effects of co-pyrolysis
lulosic biomass (Chen et al., 2017b). Consequently, Chen et al. (2019a) could be attributed to the formation of H and OH radicals, materials with
pyrolyzed amino acids with bamboo waste and reported that the in­ high hydrogen content acted as hydrogen donors. Those radicals shifted
clusion of lignocellulosic biomass into microalgae increased phenols and from biomass to char structure and led to the decomposition of char
oxygenated compounds significantly while cutting down nitrogenous (Meng et al., 2016). Duan et al. (2015) pyrolyzed waste rubber tires with
compounds sharply. The large number of –NH2 groups in proteins was Chlorella pyrenoidosa in supercritical ethanol and observed positive
the main reason for the increment of phenols, which provided numerous synergistic effects between two raw materials, which favored deni­
H+, then facilitated the scission of lignin fragments via β-O-4 linkage trogenation and deoxygenation to ameliorate the quality of bio-oil.
breaking reaction. Additionally, the Maillard reaction and ketonization Microalgae could form solid material then acted as free radical donors
reaction caused the growth of oxygenated compounds in the bio-oil, to induce the polyisoprene and poly chain scission. The hydrogen
− C=O groups from bamboo waste reacted with − NH2 groups easily. transferred from polyisoprene and poly chain to microalgae-derived
After − NH2 groups from amino acids were removed, oxygenated com­ radicals, which stabilized the primary product of microalgae thermal
pounds were formed, while the Maillard reaction produced new degradation. Microalgae-derived radicals took part in the termination of
nitrogenous compounds (Moens et al., 2004). Owing to the large amount polymer radicals to produce bio-oil, biochar, and biogas.
of NH3 released during the co-pyrolysis, amino acids generated
numerous intermediates which contain –COOH group, then reacted with 3.2.2. Synergistic effects on the bio-oil
each other to produce ketones via ketonization reaction directly As mentioned above, co-pyrolysis exhibits a positive effect on the
(Resasco & Crossley, 2015). As the reason for the decline of nitrogenous formation of high-grade bio-oil, synergistic effects are the primary fac­
compounds, numerous − NH2 groups were released out as NH3, while tor. Meanwhile, the performance of pyrolysis and the properties of bio-
some nitrogenous compounds concentrated in the biochar as the forms oil are prone to be influenced by various factors, which have been fully
of pyrrolic-N, pyridinic-N, and quaternary-N. discussed in the literature (Abnisa & Wan Daud, 2014; Azizi et al.,
Moreover, Chen et al. (2017a) conducted the co-pyrolysis of bamboo 2018a; Fakayode et al., 2020). Thus, this section mainly introduces the
with Nannochloropsis sp. and observed a sharp growth of long-chain fatty influence of synergistic effects on the yield and quality of bio-oil. As
acids and a sharp decline of acetic acids in the bio-oil. The acidic unit summarized in Table 2, synergistic effects favor the improvement of the
(− COOH) in hemicellulose of bamboo and basic unit (− NH2) in the bio-oil properties during the co-pyrolysis process. The utilization of raw
proteins of microalgae induced condensation reaction during the co- materials with high volatile matter content is instrumental in the gen­
pyrolysis process, which inhibited the formation of acetic acids. Simul­ eration of bio-oil, whereas the application of co-feedstocks with high ash
taneously, co-pyrolysis inhibited the generation of acetic acids through content hinders the process, but promotes the production of biochar and
violent decarboxylation reactions by releasing a large volume of CO2. As biogas (Dhyani & Bhaskar, 2018). Table 1 indicates plastic materials like
for the increase of long-chain fatty acids, the cracking of lipids in PP and polyvinyl chloride (PVC) possess the high volatile matter of

5
G. Su et al. Bioresource Technology 344 (2022) 126096

Table 3
Synergistic effects on the yield and HHV of bio-oil.
Microalgae Co-feedstock T(◦ C) Catalyst Bio-oil yield (wt%) HHV (MJ/kg) Reference
MA CO 1:1E 1:1D MA CO 1:1E 1:1D

Nannochloropsis sp. Bamboo 600 ◦ C − 62.50 61.00 66.63 4.88 – – – – (Chen et al., 2017a)
Nannochloropsis sp. Bamboo waste 600 ◦ C − 53.00 42.00 51.00 3.50 27.52 20.03 24.19 0.42 (Chen et al., 2018c)
Nannochloropsis sp. Bamboo waste 600 ◦ C Biochar 53.00 42.00 35.00 − 12.50 27.52 20.03 29.72 5.95 (Chen et al., 2018c)
Spirulina platensis Bamboo waste 600 ◦ C − 55.00 42.00 52.00 3.50 21.44 20.03 20.17 − 0.56 (Chen et al., 2018c)
Spirulina platensis Bamboo waste 600 ◦ C Biochar 55.00 42.00 37.00 − 11.50 21.44 20.03 25.10 4.37 (Chen et al., 2018c)
Chlorella vulgaris Rural solid waste 500 ◦ C − 53.77 48.88 47.23 − 4.10 23.80 32.20 32.18 4.18 (Tang et al., 2020a)
Chlorella vulgaris Rural solid waste 500 ◦ C CaO 50.29 47.50 48.01 − 0.88 26.60 33.60 35.80 5.70 (Tang et al., 2020a)
Chlorella vulgaris Rural solid waste 500 ◦ C MgO 49.25 46.17 48.47 0.76 25.20 31.80 30.08 1.58 (Tang et al., 2020a)
Chlorella vulgaris Rural solid waste 500 ◦ C HZSM-5 53.33 45.83 46.32 − 3.26 27.20 30.60 32.92 4.02 (Tang et al., 2020a)
Chlorella vulgaris Kitchen waste 600 ◦ C − 55.00 56.00 30.00 − 25.50 30.30 12.00 33.90 12.75 (Chen et al., 2018a)
Chlorella vulgaris PVC 800 W − 40.80 16.20 25.20 − 3.30 32.71 36.95 37.11 2.28 (Dai et al., 2018b)
Chlorella vulgaris PVC 1 kW − 50.60 24.40 41.20 3.70 33.54 30.95 37.46 5.22 (Dai et al., 2018b)
Chlorella vulgaris Waste glycerol 900 ◦ C − 56.90 73.70 58.50 − 6.80 29.50 20.03 24.25 − 0.52 (Wang et al., 2019a)
Chlorella vulgaris Wood sawdust 1 kW SiC 49.30 20.20 44.40 9.65 23.61 28.66 17.97 − 8.17 (Chen et al., 2019b)
Chlorella vulgaris Wood sawdust 1 kW AC 45.30 21.20 37.10 3.85 19.93 29.30 15.02 − 9.60 (Chen et al., 2019b)
C. pyrenoidosa Waste tire 330 ◦ C − 52.50 25.30 53.60 14.70 33.68 42.90 38.96 0.67 (Duan et al., 2015)
Chlorella Tire (Under N2) 1 kW AC 53.60 22.80 36.20 − 2.00 – – – – (Fang et al., 2018)
Chlorella Tire (Under CO2) 1 kW AC 50.20 20.40 32.60 − 2.70 – – – – (Fang et al., 2018)
Isochrysis sp. Sewage sludge 500 ◦ C − 42.60 25.20 34.37 0.47 38.00 36.50 37.90 0.65 (Wang et al., 2016)

MA: Microalgae; CO: Co-feedstock; E: Experimental value; D: Difference between experimental and theoretical value.

99.60 and 93.77 wt%, while the ash content is close to zero, followed by 2018b).
lignocellulosic biomass. While sewage sludge has the highest ash con­ EHI = (H − 2O − 2S − 3N)/C (1)
tent and coal possesses the lowest volatile matter. Therefore, co- The H, O, S, N, and C refer to the moles of hydrogen, oxygen, sulfur,
pyrolysis of microalgae with plastics or lignocellulosic biomass can nitrogen, and carbon.
facilitate the production of bio-oil, whereas the usage of coal or sewage Furthermore, HHV is another crucial criterion for the quality of bio-
sludge as a co-feedstock is difficult to raise the yield of bio-oil. oil. The high contents of carbon and hydrogen will produce bio-oil with
Some studies have reported that synergistic effects are beneficial to high HHV, whereas the high contents of oxygen and nitrogen have an
increase the yield of bio-oil. Chen et al. (2017a) performed co-pyrolysis inverse effect (Ahmed et al., 2020). Many researchers have demon­
of Nannochloropsis sp. with bamboo. The yield of bio-oil went up to strated that synergistic effects can positively influence the quality of bio-
66.63 wt% using 50 wt% of bamboo in the mixture at 600 ◦ C, while the oil by decreasing the content of oxygenated and nitrogenous compounds
bio-oil generated from the sole pyrolysis of bamboo and microalgae and carboxylic acids, and increasing the HHV. In general, the addition of
were just 60.05 and 61.85 wt%, respectively. Duan et al. (2015) ob­ co-feedstocks with high EHI is favorable to improve the quality of bio-
tained a similar conclusion via the co-pyrolysis of Chlorella pyrenoidosa oil. Xie et al. (2015) pyrolyzed Nannochloropsis sp. with scum and
and waste rubber tires. The yield of bio-oil from the pyrolysis of pure detected the positive synergistic effects on the yield and aromatics
microalgae was only 52.8 wt%, which was twice as high as that of waste proportion of bio-oil when the ratio of Nannochloropsis sp. to scum was
rubber tires. Nevertheless, the highest value of 57.2 wt% was reached less than 2:1. As a hydrogen donor, scum raised the EHI of the mixture to
through the co-pyrolysis of microalgae with 20 wt% of waste rubber boost the formation of aromatics. Simultaneously, oxygenated com­
tires. This was because co-pyrolysis inhibited the secondary cracking of pounds in the volatiles from the pyrolysis of Nannochloropsis sp. pro­
pyrolysis volatiles, thereby limiting the generation of biogas. moted the cracking of scum, further enhanced the yield and quality of
Nevertheless, some researchers have pointed out that the synergistic bio-oil. Moreover, Tang et al. (2019) conducted the co-pyrolysis of
effects can hinder the formation of bio-oil. Tang et al. (2020a) reported Nannochloropsis sp. with LDPE and found significant synergistic effects.
that the bio-oil yield declined during the co-pyrolysis of Chlorella vulgaris The growth of the LDPE ratio reduced the content of oxygenated and
and rural solid waste, only 47.23 wt% bio-oil was obtained, which was nitrogenous compounds, while prominently rose the aliphatic hydro­
lesser than each sole pyrolysis. Meanwhile, an analogous result was carbons content in the bio-oil. Oxygenates tended to evolve as H2O and
gained by Chen et al. (2018a), via the co-pyrolysis of Chlorella vulgaris CO, while nitrogenous compounds were prone to be transformed into
and kitchen waste. The highest bio-oil yield of 56.00 wt% was gained by gas products. Meanwhile, the addition of biomass with high EHI facili­
sole pyrolysis of kitchen waste, while only 30.00 wt% bio-oil was ob­ tated the quality of bio-oil was also confirmed by Fang et al. (2018) and
tained by using 50 wt% kitchen waste. The main reason was the sec­ Qi et al. (2018).
ondary reaction, which converted part of bio-oil into biogas. Moreover, Furthermore, synergistic effects positively influenced the HHV of
after mixing, feedstocks pertained high H/C ratio, and increased the bio-oil while reducing the yield of bio-oil. Chen et al. (2018a) saw a
hydrogen-related chemical groups significantly, which was instrumental decline in the bio-oil yield during the co-pyrolysis of kitchen waste and
in the formation of biogas too. Chlorella vulgaris. The synergistic effects between different materials
In addition, based on the perspective of the application, hydrocar­ increased the hydrocarbons content and decreased the carboxylic acids
bons are the most significant chemical composition of biofuels, bio-oil in the bio-oil, resulted in the growth of HHV to 33.9 MJ/kg. The same
with high content of hydrocarbons can be used as transport fuel addi­ phenomenon that synergistic effects ameliorated the bio-oil quality and
tives and industrial chemicals directly (Fan et al., 2020; Liu et al., reduced the bio-oil yield was also witnessed by Tang et al. (2020a), Dai
2020b). The high content of oxygenated compounds in bio-oil possesses et al. (2018b), and Chen et al. (2018c). Nevertheless, synergistic effects
a deleterious effect on the gasoline engine, which will cause several sometimes adversely affect the quality of bio-oil. According to the
problems owing to high acidity. Consequently, it is necessary to increase research conducted by Chen et al. (2019b), the bio-oil HHV was 29.30
the content of hydrocarbons and reduce the content of oxygenated MJ/kg and 19.93 MJ/kg from the pyrolysis of pure wood sawdust and
compounds to produce high-quality bio-oil (Mei et al., 2020). Conse­ Chlorella vulgaris with activated carbon, respectively. However, the HHV
quently, EHI has been introduced to assess the relative hydrogen content of bio-oil originated from the co-pyrolysis fell to the bottom of 15.02
of different samples as shown in Eq. (1) (Bertero et al., 2019; Dai et al., MJ/kg, but the yield of bio-oil increased owing to the high volatile

6
G. Su et al. Bioresource Technology 344 (2022) 126096

value, respectively. If YD greater than 0, suggesting positive synergistic


effects on pyrolysis; while if YD less than 0, indicating negative syner­
gistic effects on pyrolysis (Yao et al., 2017).
Table 3 summarizes the synergistic effects of co-pyrolysis of micro­
algae with other biomass wastes on the yield and HHV of bio-oil. Syn­
ergistic effects have a complex effect on the bio-oil yield but induce a
positive influence on the bio-oil HHV in most cases. Fig. 1 depicts the
influence of synergistic effects originated from the co-pyrolysis of
microalgae with some representative materials on the bio-oil properties,
the accession of plastics and tires presented great strengths, which
increased the yield and HHV of bio-oil by 3.7 wt% and 5.22 MJ/kg, 14.7
wt% and 0.67 MJ/kg, respectively. The inclusion of bamboo waste also
brought a decent result, increased the yield and HHV of bio-oil by 3.50
wt% and 0.42 MJ/kg, while the usage of sewage sludge had a tiny
enhancement on the yield and HHV of bio-oil. Other materials also
exhibited impressive performances, the addition of rural solid waste and
kitchen waste enhanced the bio-oil HHV, whereas the application of
wood sawdust raised the bio-oil yield without the amelioration on the
bio-oil HHV due to its high volatile matter content and low EHI. In short,
tires, plastics, and bamboo waste are suitable co-feedstocks for the py­
rolysis of microalgae.
Fig. 2 presents the chemical composition of bio-oil generated from
the co-pyrolysis of Chlorella vulgaris with other materials. The addition of
Fig. 1. Synergistic effects on the yield and HHV of bio-oil (Chen et al., 2018a;
Chen et al., 2019b; Chen et al., 2018c; Dai et al., 2018b; Duan et al., 2015; Tang
50 wt% different raw materials possessed a dramatic effect on the
et al., 2020a; Wang et al., 2019a; Wang et al., 2016). *NS: Nannochloropsis sp.; chemical composition of bio-oil. At first, there was a remarkable
CP: C. pyrenoidosa; CV: Chlorella vulgaris; IS: Isochrysis sp.; PVC: Polyvi­ reduction of nitrogenous compounds after adding each material, as
nyl chloride. numerous nitrogenous compounds were released into the air or
concentrated in the biochar. Additionally, the percentage of hydrocar­
bons increased significantly during the co-pyrolysis of microalgae with
kitchen waste and wood sawdust. Synergistic effects between those
materials promoted aromatization and deoxygenation reactions. How­
ever, the addition of waste glycerol reduced the content of hydrocarbons
sharply with the prominent increment in the content of alcohols. This
was ascribed to that glycerol was pyrolyzed into ethylene glycol, then
further dehydrated to form alcohols. Moreover, additional ester com­
pounds were generated with the drop of ketones and aldehydes during
the co-pyrolysis process (Wang et al., 2019a).
In brief, the addition of co-feedstocks with high volatile matter
content is conducive to the formation of bio-oil. While the application of
materials with high EHI can induce a series of reactions, such as Diels-
Alder, dehydration, hydrogen transfer, Maillard, ketonization, conden­
sation, denitrogenation, aromatization, and deoxygenation reactions,
which inhibits the generation of oxygenated and nitrogenous com­
pounds and promotes the production of aromatic hydrocarbons, thereby
improving the quality of bio-oil. Simultaneously, as a substitute for fossil
fuel, bio-oil is the most economically and environmentally important
product. The large-scale production of bio-oil is the only way to achieve
Fig. 2. Chemical composition of bio-oil from the co-pyrolysis of Chlorella vul­
garis (CV) with other biomass wastes (Chen et al., 2019b; Wang et al., 2019a; the commercialization and industrialization of related industries.
Yuan et al., 2015). Consequently, it is particularly important to search and develop excel­
lent and low-cost co-feedstocks. Many materials with high volatile
matter content of two raw materials. Nevertheless, the phenomenon was matter content and EHI, such as medical waste, waste cooking oil, and
waste lubricating oil, need to be studied to further ameliorate the
not common during the co-pyrolysis process.
Synergistic effects can be realized through the interaction of free properties of bio-oil from the pyrolysis of microalgae.
radicals, while the characteristics of various feedstocks and their
blending ratio are the primary factors influencing the synergistic effects 4. Catalytic effects on the co-pyrolysis process
and resulting in complex and varied pyrolysis (Wang et al., 2019b; Zhao
et al., 2020). Usually, the difference between the theoretical and Catalytic pyrolysis is a prospective technique to facilitate the existing
experimental results of co-pyrolysis is a vital criterion to determine the co-pyrolysis process and the properties of bio-oil. Furthermore, the
synergistic effects. application of catalysts is conducive to lower the activation energy,
YT = W1X1 + W2X2 (2) reduce acidity, shorten the reaction time, and affect the product distri­
YD = YE- YT (3) butions (Lee et al., 2020a). The catalytic pyrolysis process is categorized
As shown in Eq. (2) and Eq. (3), W1 and W2 represent the mass as in-situ and ex-situ modes. For the in-situ process, raw materials and
proportion of each raw material in the mixture samples, respectively. X1 catalysts are mixed before co-pyrolysis, while for the ex-situ process, the
and X2 refer to the product or physic-chemical properties of every single pyrolysis volatiles pass through the catalyst at an operational tempera­
raw material, YE and YT are the experimental value and the theoretical ture (Yu et al., 2019). Table 4 summarizes catalytic effects on the co-
pyrolysis of microalgae and other biomass wastes. Numerous studies

7
G. Su et al. Bioresource Technology 344 (2022) 126096

Table 4
Catalytic effects on the co-pyrolysis of microalgae and other biomass wastes.
Microalgae Co- Catalyst Conditions Reactor Catalyst performance Reference
feedstock

Chlorella vulgaris Rural solid CaO 500 ◦ C, 25 Fixed-bed • Increases the activation energy and HHV. (Tang et al.,
waste min, N2 reactor • Raises the content of aliphatic hydrocarbons. 2020a)
• Reduces acids and nitrogenous compounds.
Chlorella vulgaris Rural solid HZSM-5 500 ◦ C, 25 Fixed-bed • Reduces the activation energy. (Tang et al.,
waste min, N2 reactor • Increases the aromatic hydrocarbons and nitrogenous 2020a)
compounds content in the bio-oil.
Chlorella vulgaris Rural solid MgO 500 ◦ C, 25 Fixed-bed • Increases the hydrocarbons content in the bio-oil. (Tang et al.,
waste min, N2 reactor • Reduces the activation energy. 2020a)
Nannochloropsis sp. Palm kernel Cu/HZSM-5 600 ◦ C, 15 Fixed-bed • Increases the aromatics content in the bio-oil. (Chang et al.,
shell min, N2 reactor • Cu reduces the number of strong acid sites of HZSM-5. 2018)
Nannochloropsis sp. Palm kernel CaO/HZSM- 600 ◦ C, 15 Fixed-bed • Increases the aromatics content in the bio-oil. (Chang et al.,
shell 5 min, N2 reactor • Generates lower peak area values of specific aromatics than 2018)
other catalysts.
Chlorella vulgaris Kitchen CaO 900 ◦ C, 40 ◦ C/ Fixed-bed • Decreases the acids content sharply. (Chen et al.,
waste min, N2 reactor • Increases the aromatic hydrocarbons content. 2018b)
Chlorella vulgaris Kitchen CaCO3 900 ◦ C, 40 ◦ C/ Fixed-bed • Decreases the acids content. (Chen et al.,
waste min, N2 reactor • Improves the benzene, toluene, ethylbenzene, xylenes for all 2018b)
samples and hydrocarbons in the bio-oil for sole microalgae
and mixture, except kitchen waste.
Chlorella vulgaris Kitchen Permutit 900 ◦ C, 40 ◦ C/ Fixed-bed • Lowers the initial temperature. (Chen et al.,
waste min, N2 reactor • Results in less residues. 2018b)
• Shows unfavorable selectivity on the products.
Chlorella vulgaris Kitchen SiO2 900 ◦ C, 40 ◦ C/ Fixed-bed • Shows an unsatisfied product distribution. (Chen et al.,
waste min, N2 reactor • Hinders the formation of hydrocarbons and benzene, 2018b)
toluene, ethylbenzene, xylenes.
• Promotes the generation of oxygenated and nitrogenous
compounds.
Chlorella vulgaris Wood Activated 1 kW, 25 min, Microwave • Increases the gas yield and aromatic hydrocarbons content in (Chen et al.,
sawdust carbon N2 reactor the bio-oil. 2019b)
• Possesses a better absorptive capacity of microwave
irradiation.
Chlorella vulgaris Wood SiC 1 kW, 25 min, Microwave • Raises the bio-oil and biochar yield. (Chen et al.,
sawdust N2 reactor • Increases the content of aliphatic hydrocarbons and phenols 2019b)
in the bio-oil.
Nannochloropsis sp. Scum HZSM-5 550 ◦ C, 30 Microwave • Increases the gas yield and hydrocarbons content. (Xie et al.,
min, He reactor • Decreases oxygenated and nitrogenous compounds in the 2015)
bio-oil.
Spirulina platensis, Bamboo Biochar 600 ◦ C,30 min, Two-stage fixed- • Facilitates the decomposition of long-chain fatty acids and (Chen et al.,
Nannochloropsis sp. waste Ar bed reactor O− species in the bio-oil. 2018c)
• Enhances the production of phenols and aromatics in the bio-
oil.
• Reduces oxygenated compounds content in the bio-oil.

have reported that the presence of catalysts has a positive effect on the hydrocarbons from 43.38 to 93.29 wt%, a notable drop in the content of
pyrolysis process and product properties. olefins from 33.19 to 2.29 wt% and a severe decline in the content of
nitrogenous compounds from 18.37 to 0.59 wt%, when the mass ratio of
catalyst/feedstock raised from 1 to 10. They elucidated the phenomenon
4.1. Catalytic effect of zeolites that volatiles fully reached the surface of HZSM-5, the intermediate
polar product was produced by degradation during the pyrolysis of
At present, over 60 natural zeolites and 230 zeotype frameworks Nannochloropsis sp., the nonpolar compounds and hydrocarbons were
with special characteristics and porous structures have been discovered formed by the destruction of nitrogenous functional groups. Addition­
and designed. There are various kinds of zeolites including ZSM-5, ally, the conclusion of HZSM-5 can greatly decrease the nitrogenous
HBEA, HZSM-5, and permutit, which have been utilized in the co- compounds and improve aromatics selectivity in the bio-oil was also
pyrolysis of microalgae and other biomass wastes. Carbohydrates, al­ supported by Zhang et al. (2017) and Yang et al. (2011). On top of that,
cohols, acids, aldehydes, furans, ketones, and other functional groups Xie et al. (2015) studied the effect of HZSM-5 on the properties of bio-oil
can easily diffuse into the zeolite molecular sieve channels, participate during the co-pyrolysis of Nannochloropsis sp. and scum. The bio-oil
in a series of reactions to produce oxygen, aromatics, and coke (Zhang yield declined with the presence of catalyst as the thermal cracking re­
et al., 2016). action converted the condensable gas into non-condensable gas. Mean­
As a consequence of the existence of proteins and high moisture while, the content of polycyclic aromatic hydrocarbons and aromatic
content, the bio-oil originated from the pyrolysis of microalgae possesses hydrocarbons raise sharply, while the content of aliphatic hydrocarbons,
high oxygen and nitrogen contents. Therefore, HZSM-5 has been widely oxygenated, and nitrogenous compounds declined. Due to its unique
used owing to its characteristics of deoxygenation, denitrogenation, and structure, HZSM-5 manifested excellent shape-selectivity to the aro­
aromatization. HZSM-5 is manufactured from ZSM-5 via calcined and matics and drove the cracking and deoxidation of oxygenated com­
hydrothermal treatments and manifests strong acidity and favorable pounds. Furthermore, a similar conclusion was also gained in the
deoxygenation effect because of active hydrogen, which can facilitate pyrolysis of lignocellulosic biomass (Mihalcik et al., 2011).
the formation of aromatic hydrocarbons and hinder the generation of However, many studies have revealed that the HZSM-5 can poly­
oxygenated compounds like acids, alcohols, and ketones (State et al., merize a considerable number of oxygenated compounds to form coke
2019). Qi et al. (2018) conducted the co-pyrolysis of PP with Nanno­ on the surface of the catalyst, thus reduces its catalytic activity (Wu
chloropsis sp. and witnessed a sharp growth in the content of aromatic

8
G. Su et al. Bioresource Technology 344 (2022) 126096

et al., 2015). To overcome the barrier, the utilization of modified HZSM- acids. While the utilization of CaCO3 did not improve the hydrocarbons
5 has attracted attention rapidly. The chemical liquid deposition tech­ content in the bio-oil from the pyrolysis of pure kitchen waste, the
nique is applied to ameliorating the catalytic performance of HZSM-5 by addition of SiO2 presented an unfavorable product distribution, and the
narrowing the pore openings and reducing the number of external acid presence of permutit did not display any obviously admirable selectivity
sites, which has advantages of simple apparatus, easy operation, and low on the final products.
cost (Dai et al., 2018a). Chang et al. (2018) performed the co-pyrolysis Tang et al. (2020a) carried out the co-pyrolysis of Chlorella vulgaris
of Nannochloropsis sp. and palm kernel shell with HZSM-5, CaO/HZSM- and rural solid waste with the aid of three kinds of catalysts (HZSM-5,
5, and modified HZSM-5 with Cu. These three types of catalysts raised MgO, and CaO) and detected that non-catalytic co-pyrolysis had an
the aromatics content from 30.39 to 40.19, 36.12, and 42.16 wt%, inhibitory influence on the bio-oil yield, primarily related to the sec­
respectively. The results revealed that Cu/HZSM-5 was the best catalyst ondary reaction to form biogas. While the addition of catalysts allevi­
which enhanced the quality of bio-oil effectively, as Cu cut down the ated the situation, the presence of MgO even reversed the trend and
number of strong acid sites on the surface of HZSM-5 and improved the generated positive synergistic effects on the bio-oil yield. As for the
catalytic activity, while pore structures of HZSM-5 facilitated the for­ HHV, synergistic effects influenced the HHV of bio-oil positively, CaO
mation of aromatics. In addition, HZSM-5 has displayed great perfor­ further enhanced the tendency due to its strong deoxygenation, deni­
mance in the reduction of activation energy during the co-pyrolysis trogenation, and aromatization capacity. Based on the above analysis,
process. Bong et al. (2020) and Dai et al. (2019) reported that the CaO showed the best catalytic performance based on the amelioration of
average activation energy fell from 142.56 to 131.37 kJ/mol and 258.72 bio-oil yield and quality.
to 198.66 kJ/mol using the Flynn-Wall-Ozawa model with the help of
HZSM-5 for co-pyrolysis of Chlorella vulgaris with peanut shell and the 4.3. Catalytic effect of carbon-based catalysts
co-pyrolysis of Spirulina with oil shale, respectively. Furthermore, Tang
et al. (2020a) obtained the same result that the co-pyrolysis of Chlorella Carbon-based catalysts also play critical roles in the co-pyrolysis of
vulgaris and rural solid waste using HZSM-5 catalyst has decreased the microalgae and other biomass wastes. Carbon-based catalysts are effi­
activation energy from 258.33 to 222.64 kJ/mol via Kissinger-Akahira- cient catalysts owing to the carbonaceous matter and unique surface
Sunose method. area, while mineral content (Al2O3, MgO, and Fe2O3) and metal (Na, Mg,
Except for HZSM-5 and its derivatives, other zeolite was also K, Al, Ca, and P) can further facilitate its catalytic activity (Gao et al.,
employed in the co-pyrolysis but did not achieve ideal results. Chen et al. 2020). Furthermore, carbon-based catalysts such as biochar, SiC, acti­
(2018b) pyrolyzed the Chlorella vulgaris and kitchen waste with the aid vated carbon, and graphite are widely utilized in microwave-assisted
of permutit and ascertained that the catalyst had a predominant pro­ pyrolysis as absorbents with admirable microwave absorptivity.
moting effect on the reduction of initial temperature and residue mass, Numerous studies have demonstrated that biochar possesses strong
but did not show positive selectivity on the final products. nitrogen fixation ability and deoxidation power. Chen et al. (2018d)
pyrolyzed Nannochloropsis sp. and cellulose under the aid of biochar.
4.2. Catalytic effect of metal oxides They ascertained that HCN, NH3, and nitrogenous intermediates had
reactions with high active oxygen species, promoted the formation of
Metal oxides exhibit Lewis acid and Bronsted base because of the pyrrolic-N, pyridinic-N, amine/amide-N, and quaternary-N on the sur­
positive metal ions and negative oxygen ions. They have been explored face of the catalyst, which made the high concentration of nitrogen in
excessively to improve the properties of bio-oil with the characteristics the char and the reduction of N-containing gas, amines, and nitriles.
of acid base and redox. CaO, MgO, CaCO3, ZnO, and SiO2 are the most Chen et al. (2018c) noticed that the usage of biochar in the co-pyrolysis
commercial metal oxides applied in the co-pyrolysis of microalgae and of bamboo waste and Spirulina platensis had a huge impact on the
other biomass wastes, which can be divided into basic metal oxides product distribution, which reduced the bio-oil yield from 52 to 37 wt%,
(CaO, MgO) and acidic metal oxides (SiO2) (Chan & Tanksale, 2014). while increasing the biogas yield from 12 to 20 wt%. Simultaneously,
Generally, the basic metal oxides are suitable for aldol condensation and the addition of biochar enhanced the content of phenol and long-chain
ketonization of carbonyl compounds and carboxylic acid, while the fatty acids and hindered the formation of oxygenated and nitrogenous
acidic metal oxides present better catalytic activity during the deoxy­ compounds in the bio-oil, rose the bio-oil HHV to 25.10 MJ/kg.
genation process. Furthermore, other carbon-based additives also exhibited great perfor­
Microalgae possess a high content of oxygen and nitrogen, thus the mance on denitrogenation and deoxygenation. Chen et al. (2019b) re­
catalyst with the ability of denitrogenation and deoxygenation is ported the carbon-based additives improved the quality of bio-oil by
essential for the pyrolysis of microalgae. Among metal oxides, CaO deoxygenation and aromatization through the microwave-assisted co-
manifests the best performance on the reduction of acids and nitrides pyrolysis of Chlorella vulgaris with wood sawdust. The presence of SiC
and the improvement of hydrocarbons in the bio-oil. Tang et al. (2020a) raised the yield of bio-oil and biochar, while the application of activated
carried out the co-pyrolysis of Chlorella vulgaris and rural solid waste carbon increased the biogas yield due to the good absorptive capacity of
with the assistance of three different catalysts. CaO raised the activation microwave irradiation. Moreover, both additives had a positive impact
energy of the overall reaction, while HZSM-5 and MgO diminished it. on the aromatization and deoxygenation by increasing the content of
Furthermore, CaO increased aliphatic hydrocarbons content especially aliphatic hydrocarbons and phenols in the bio-oil, but denitrogenation
light fraction, reduced acids and nitrogenous compounds content. MgO was not detected during the co-pyrolysis process.
boosted the increase of hydrocarbons and the decrease of carboxylic In summary, the addition of CaO and Cu/HZSM-5 exhibits great
acids and nitrogenous compounds for sole pyrolysis of microalgae but advantages in previous researches, which can facilitate the formation of
failed to the pyrolysis of rural solid waste and all mixtures. Moreover, aromatics and inhibit the generation of oxygenated and nitrogenous
HZSM-5 raised aromatic hydrocarbons content, but unfortunately, also compounds. The overall bio-oil quality has been improved effectually
increased nitrogenous compounds content. The analogous experimental because Ca2+ in CaO neutralizes acids. Meanwhile, the pore structure
result was obtained by Chen et al. (2018b), via the co-pyrolysis of and surface acidity of HZSM-5 possess strong deoxygenation, deni­
Chlorella vulgaris and kitchen waste with the aid of CaCO3, CaO, SiO2, trogenation, and aromatization capabilities, the usage of Cu cuts down
and permutit. The process was favored by CaO without exception, the the number of strong acid sites of the catalyst. Modified HZSM-5 hinders
content of hydrocarbons in the bio-oil increased for all mixed samples. the formation of coke and ameliorates the activity of catalysts, which
Moreover, the acids yield decreased sharply by 2.4 times for pure can be a prospective research direction for the co-pyrolysis of micro­
kitchen waste, 6.1 times for pure microalgae, and 4.5 times for the algae and other materials. Additionally, some studies have proven the
mixture, the metal cation Ca2+ from CaO was capable to neutralize effectiveness of acid and alkali coexist bifunctional catalyst in the

9
G. Su et al. Bioresource Technology 344 (2022) 126096

reduction of acid content, the inhibition of coke formation, and the to 18.7%. Jonker and Faaij (2013) treated microalgae as feedstocks for
growth of bio-oil yield. Yu et al. (2020) utilized coexist bifunctional the production of bioenergy and found the production cost of biofuels
catalyst based on SBA-15 to conduct the pyrolysis of waste cooking oil was 136 €/GJ for raceway ponds and 153 €/GJ for horizontal tubular
and achieved satisfying results. Microalgae possess the drawbacks of systems, which was far more than diesel and gasoline (5–20 €/GJ), even
high oxygen and ash content. Hence, the acid and alkali coexist the costs could be reduced by a quarter via possible improvement op­
bifunctional catalyst can be explored in the microalgae co-pyrolysis. tions during the production process. Nevertheless, co-pyrolysis with
Moreover, the relationship between the product properties and cata­ low-cost raw materials and co-production of bioenergy and high-value
lysts still requires additional studies, whereas the suitable catalyst for products has not been taken into consideration.
specific microalgae and other biomass wastes needs to be further The techno-economic assessment for co-pyrolysis of microalgae and
developed. other biomass wastes was conducted by Wang et al. (2016) and Chen
et al. (2018c). Wang et al. (2016) detected that the inclusion of Isochrysis
5. Economic viability of the co-pyrolysis process sp. cut down the energy consumption and facilitated energy efficiency.
Co-pyrolysis saved energy costs and improved the energy conversion
Many studies have shown the great techno-economic strengths of co- and stability of sewage sludge pyrolysis. Moreover, Chen et al. (2018c)
pyrolysis in comparison with sole pyrolysis. The equipment of co- performed an economic assessment for the mass production of biofuels
pyrolysis is similar to sole pyrolysis, which shows simplicity in design via the co-pyrolysis of microalgae and bamboo waste. However, co-
and operation, but the result is considerably profitable owing to the pyrolysis did not show great advantages, sole pyrolysis of Spirulina
reduction of feedstocks cost and the improvement of products properties platensis earned the most profit of 33.6 million RMB annually, followed
(Ozyuguran et al., 2018). by sole pyrolysis of Nannochloropsis sp. with 32.8 million RMB every
Meyer et al. (2020) conducted the economic and environmental year, while the third one was the co-pyrolysis of Spirulina platensis and
benefits analysis for the conversion of 11 biomass feedstocks to biofuels bamboo waste with 31.6 million RMB per year. However, the input and
through fast pyrolysis or co-pyrolysis. The study concluded that capital- output were only taken into account, while the positive social, eco­
related costs (including capital depreciation, return on investment, and nomic, and environmental effects owing to the disposal of bamboo waste
income tax) was the maximum cost contributed with 30–40% of the were not considered in the study. In addition, anther raw materials with
minimum fuel selling price, followed by feedstock cost (30%), hydro­ low cost would push the co-pyrolysis process forward.
treating catalyst cost (13–18%), and labor cost (12–15%). Woody and
herbal mixed raw materials with reasonable product yield, little risk, 6. Challenges and prospects of the co-pyrolysis process
low raw materials price, and low transportation costs display tremen­
dous advantages over single raw materials. Kuppens et al. (2010) As mentioned above, the economic feasibility of microalgae-based
analyzed the profitability of the co-pyrolysis of biopolymers with willow biodiesel is still poor due to the high cost of microalgae and low prod­
and indicated that co-pyrolysis can obtain better economic benefits in uct benefits. However, the cost of microalgae is expected to decline
comparison with sole pyrolysis of willow, the net present value of cash significantly in the foreseeable future with the support of government
flow raised by at least 0.405 MEUR because of the growth of high-value policies and various capitals. For instance, the Chinese government has
crotonic acid. Zhang et al. (2015) pyrolyzed corn stalk and HDPE in the initiated many national key research and development projects for
presence of HZSM-5 and performed an economic assessment. Value microalgae biomass and biofuels to regulate and incentivize the devel­
index formula was used to analyze the economic potential of aromatic opment of the microalgae industry. In 2009, the “Key technologies in
hydrocarbons and found that high content of aromatic hydrocarbons CO2 to oil algae to biodiesel” project received 3.2 million US dollars
was achieved when the hydrogen to carbon effective ratio of feedstock funding from the Chinese government. In 2013, ENN Group got a loan of
was more than 1. Co-pyrolysis was superior to sole pyrolysis because of 46.5 million US dollars from the China Development Bank to scale up
high-quality products and subsequent economic benefits. the microalgae farm to 280 Hectares in Inner Mongolia (Chen et al.,
As for the production of microalgae, the high cost of cultivation and 2020). The U.S. government has launched many microalgae programs to
the low production capacity of microalgae are the primary barriers to support the target of 17% of transportation fuels in the United States can
the commercialization of the microalgae production industry. Many be replaced by microalgae-based biofuels (Su et al., 2017). Meanwhile,
scholars have devoted themselves to the techno-economic assessment of in the background of global warming, carbon neutrality and low-carbon
microalgae cultivations and made a substantial number of achieve­ economy are the themes of sustainable development in the world. The
ments. Rezvani et al. (2016) took a techno-economic assessment on the microalgae industry is closely linked to the topic and will receive more
microalgae cultivations and concluded that the selling price is between attention and strong support from governments of all countries. Because
$440–1028/t when photosynthetic efficiency is 4%. Yang et al. (2019) the cultivation and utilization of microalgae can bring numerous social,
demonstrated that with the rapid development of cultivation technology economic, and environmental benefits, like carbon dioxide reduction,
last decade, the cost of biocrude of microalgae was decreased to wastewater treatment, and environmental improvement, which will
approximately $7.50 per gallon, while the low productivity of micro­ gain benefits from variegated policy arbitrations like second-generation
algae was still the main bottleneck. biomass, such as tax break, economic support, subsidies for production
A feasibility and economic analysis of microalgae-based biodiesel and infrastructure (Sajjadi et al., 2018). Additionally, many microalgae
production in China was conducted by Sun et al. (2019b). The produc­ companies have aroused the attention of commercial capital. For
tion cost of microalgae-based biodiesel was calculated as USD 2.29/kg example, Xiaozao Tech in China has completed a round of Series B
according to the 100 ha of production facilities, which was far more than funding (15.5 million US dollars) in 2021 to expand production scale
that of conventional fuel. The main obstacles were the low productivity and increase market share. Therefore, the cost of microalgae will be
of microalgae and product benefits. Additionally, Xin et al. (2016) decreased sharply in the future with the joint endeavors from various
performed a techno-economic analysis of biofuel production from the parties.
pyrolysis of wastewater-based algal. Capital, land, operation, chemical In addition, further valorization of microalgae through deep pro­
coagulation harvesting costs, and reasonable bio-oil yield and revenue cessing is vital to the microalgae industry. Microalgae possess a wide
were calculated. The estimated break-even selling price of biofuels was range of applications and enormous economic value. For instance,
USD 2.23 per gallon, which was within an acceptable range. Meanwhile, polysaccharides are used for human nourishment, cosmetics, pharma­
if breakthroughs could be made in the three key components of culti­ ceuticals, and therapeutics. Lipids in the microalgae are applied to
vation, harvesting, and downstream transformation, this method would neutraceutics, food technology, and medicine. Bioactive compounds are
obtain better overall benefits and the internal rate of return would raise extensively employed in antibiotics, agronomic chemistry, vaccines, and

10
G. Su et al. Bioresource Technology 344 (2022) 126096

Fig. 3. Flow diagram of microalgae biorefinery and biofuels production.

chemicals. Proteins are the raw material for pharmacy vaccines, nutri­ residues will be employed in the co-pyrolysis process to produce bio­
tion, and recombinant protein. Pigments have wide application in fuels. While biogas can be reused for the pretreatment process to cut
fragrance, cosmetics, and medicine (Dasan et al., 2019). Consequently, a down the storage, transportation, and heating cost, bio-oil and biochar
biorefinery system is required to realize the valorization of microalgae. are the main products for industry and transportation, biochar can also
Vanthoor-Koopmans et al. (2013) proposed an effective microalgae be utilized in the co-pyrolysis process as a catalyst. The comprehensive
biorefinery scheme, which adopted pulsed electric field as a cell facility inherits the advantages of co-pyrolysis and microalgae, which
disruption technique and ionic liquids to separate hydrophilic and hy­ will maximize the value of microalgae and accelerate the speed of
drophobic compounds. Gerardo et al. (2014) held a view that membrane commercialization of related industries (van Schalkwyk et al., 2020).
filtration was a promising technology and would play a key role in the
recovery of proteins, lipids, and carbohydrates from microalgae. 7. Conclusions
Furthermore, Banu et al. (2020) proposed three biorefinery routes
including biodiesel production, pigments, animal feed, biogas, bio- Co-pyrolysis of microalgae and other biomass wastes is an effective
hydrogen production, and concluded that process integration was way to produce high-quality biofuels. The positive synergistic effects
conductive to reduce the cost and increase the profit effectively. between various materials and the improvement of the target product
Nevertheless, present research of conventional or advanced pyrolysis of have verified the strength and effectiveness of co-pyrolysis. Simulta­
microalgae is still in the laboratory stage, large-scale production of neously, the addition of catalysts has significantly facilitated the quality
bioproduct is essential to achieve its industrialization and commercial­ of bio-oil by reducing the content of oxygenated and nitrogenous com­
ization. Therefore, the expansion of research is necessitated to complete pounds, while increasing the HHV and the content of aromatic hydro­
the industrial production and subsequent market applications of bio-oil carbons in the bio-oil. Meanwhile, a comprehensive facility that inherits
from microalgae pyrolysis. Meanwhile, the economic assessment of the superiorities of co-pyrolysis and microalgae has been proposed to
microalgae pyrolysis on a large scale is also important, because the realize the valorization of microalgae and promote the commercializa­
positive result will absorb capital and talent, thereby stimulating the tion of related industries.
development of related industries.
As the growth, survival, and reproduction organisms in aquatic CRediT authorship contribution statement
ecosystems, microalgae possess high moisture content, which is unfa­
vorable to the production of high-quality bio-oil. Accordingly, pre­ Guangcan Su: Investigation, Formal analysis, Writing – original
treatment processes, like drying or torrefaction, are imperative to draft. Hwai Chyuan Ong: Conceptualization, Writing – review & edit­
ameliorate the bio-oil quality. However, the implementation of drying ing, Supervision. Yong Yang Gan: Validation, Writing – review &
or torrefaction will substantially increase the production cost since they editing. Wei-Hsin Chen: Resources, Writing – review & editing. Cheng
are energy-intensive processes. The application of microwave-assisted Tung Chong: Formal analysis, Validation. Yong Sik Ok: Visualization,
pyrolysis can combine drying and pyrolysis processes due to the polar Writing – review & editing.
nature of water, which will simplify the equipment and reduce the cost.
It is reported that the release of water occurs between ambient tem­
perature to 200 ◦ C (Sait & Salema, 2015). Meanwhile, vacuum pyrolysis Declaration of Competing Interest
requires further exploration to increase the bio-oil yield and reduce the
production cost, because the condition of negative pressure can cut The authors declare that they have no known competing financial
down the boiling point of bio-oil, thereby promoting the generation of interests or personal relationships that could have appeared to influence
bio-oil. Lam et al. (2019) conducted the co-pyrolysis of plastics with the work reported in this paper.
waste oils using inert gas and vacuum pump, then obtained a result that
the production cost under vacuum environment was 20.47% less than References
that under N2 environment. Moreover, it is necessary to achieve excess
energy in the final product due to the endothermic characteristics of Abdullah, B., Syed Muhammad, S.A.F.a., Shokravi, Z., Ismail, S., Kassim, K.A., Mahmood,
A.N., Aziz, M.M.A. 2019. Fourth generation biofuel: A review on risks and mitigation
pyrolysis. Further studies on the selection of cheap and excellent co-
strategies. Renewable and Sustainable Energy Reviews, 107, 37-50.
feedstocks, development of effective catalysts, and mechanism of syn­ Abnisa, F., Wan Daud, W.M.A., 2014. A review on co-pyrolysis of biomass: An optional
ergistic effects behind the co-pyrolysis of microalgae and other materials technique to obtain a high-grade pyrolysis oil. Energy Conversion and Management
87, 71–85.
will contribute to accomplishing the goal.
Ahmed, Mohamed H.M., Batalha, Nuno, Mahmudul, Hasan M.D., Perkins, Greg,
As depicted in Fig. 3, the economic advantages of co-pyrolysis and Konarova, Muxina, 2020. A review on advanced catalytic co-pyrolysis of biomass
microalgae biorefinery are supposed to be combined, a comprehensive and hydrogen-rich feedstock: Insights into synergistic effect, catalyst development
facility consisted of microalgae cultivation, biorefinery, and biofuels and reaction mechanism. Bioresource Technology 310, 123457. https://doi.org/
10.1016/j.biortech.2020.123457.
production system will achieve the target. During cultivation, micro­ Ali, I., Naqvi, S.R., Bahadar, A., 2018. Kinetic analysis of Botryococcus braunii pyrolysis
algae will absorb CO2 released by vehicles using biofuels, while micro­ using model-free and model fitting methods. Fuel 214, 369–380.
algae will be used as raw materials for the production of biofuels. The Anand, V., Sunjeev, V., Vinu, R., 2016. Catalytic fast pyrolysis of Arthrospira platensis
(spirulina) algae using zeolites. Journal of Analytical and Applied Pyrolysis 118,
carbon cycle based on microalgae conforms to the principle of sustain­ 298–307.
able development and circular economy. After cultivation, microalgae Azizi, K., Keshavarz Moraveji, M., Abedini Najafabadi, H., 2018a. A review on bio-fuel
will be sent into biorefinery systems and downstream industries, the production from microalgal biomass by using pyrolysis method. Renewable and
Sustainable Energy Reviews 82, 3046–3059.

11
G. Su et al. Bioresource Technology 344 (2022) 126096

Azizi, K., Keshavarz Moraveji, M., Abedini Najafabadi, H., 2018b. Simultaneous pyrolysis Dhyani, V., Bhaskar, T., 2018. A comprehensive review on the pyrolysis of lignocellulosic
of microalgae C. vulgaris, wood and polymer: The effect of third component biomass. Renewable Energy 129, 695–716.
addition. Bioresource Technology 247, 66–72. Duan, P., Jin, B., Xu, Y., Wang, F., 2015. Co-pyrolysis of microalgae and waste rubber tire
Azizi, K., Keshavarz Moraveji, M., Arregi, A., Amutio, M., Lopez, G., Olazar, M., 2020. On in supercritical ethanol. Chemical Engineering Journal 269, 262–271.
the pyrolysis of different microalgae species in a conical spouted bed reactor: bio- EIA, 2020. Annual Energy Outlook 2020 with projections to 2050. U.S. DOE Energy
fuel yields and characterization. Bioresource Technology 123561. Information Administration, Washington, DC.
Banu, J.R., Preethi, Kavitha, S., Gunasekaran, M., Kumar, G. 2020. Microalgae based Enamala, M.K., Enamala, S., Chavali, M., Donepudi, J., Yadavalli, R., Kolapalli, B.,
biorefinery promoting circular bioeconomy-techno economic and life-cycle analysis. Aradhyula, T.V., Velpuri, J., Kuppam, C., 2018. Production of biofuels from
Bioresource Technology, 302, 122822. microalgae - A review on cultivation, harvesting, lipid extraction, and numerous
Benedek, J., Sebestyén, T.-T., Bartók, B., 2018. Evaluation of renewable energy sources applications of microalgae. Renewable and Sustainable Energy Reviews 94, 49–68.
in peripheral areas and renewable energy-based rural development. Renewable and Fakayode, O.A., Aboagarib, E.A.A., Zhou, C., Ma, H., 2020. Co-pyrolysis of
Sustainable Energy Reviews 90, 516–535. lignocellulosic and macroalgae biomasses for the production of biochar – A review.
Bertero, M., García, J.R., Falco, M., Sedran, U., 2019. Equilibrium FCC catalysts to Bioresource Technology 297, 122408.
improve liquid products from biomass pyrolysis. Renewable Energy 132, 11–18. Fan, L., Ruan, R., Li, J., Ma, L., Wang, C., Zhou, W., 2020. Aromatics production from fast
Bong, J.T., Loy, A.C.M., Chin, B.L.F., Lam, M.K., Tang, D.K.H., Lim, H.Y., Chai, Y.H., co-pyrolysis of lignin and waste cooking oil catalyzed by HZSM-5 zeolite. Applied
Yusup, S., 2020. Artificial neural network approach for co-pyrolysis of Chlorella Energy 263, 114629.
vulgaris and peanut shell binary mixtures using microalgae ash catalyst. Energy 207, Fang, S., Gu, W., Dai, M., Xu, J., Yu, Z., Lin, Y., Chen, J., Ma, X., 2018. A study on
118289. microwave-assisted fast co-pyrolysis of chlorella and tire in the N2 and CO2
Carvalho, J.C., Magalhães, A.I., de Melo Pereira, G.V., Medeiros, A.B.P., Sydney, E.B., atmospheres. Bioresource Technology 250, 821–827.
Rodrigues, C., Aulestia, D.T.M., de Souza Vandenberghe, L.P., Soccol, V.T., Soccol, C. Fermoso, J., Corbet, T., Ferrara, F., Pettinau, A., Maggio, E., Sanna, A., 2018. Synergistic
R., 2020. Microalgal biomass pretreatment for integrated processing into biofuels, effects during the co-pyrolysis and co-gasification of high volatile bituminous coal
food, and feed. Bioresource Technology 300, 122719. with microalgae. Energy Conversion and Management 164, 399–409.
Chan, F.L., Tanksale, A., 2014. Review of recent developments in Ni-based catalysts for Gan, Y.Y., Ong, H.C., Show, P.L., Ling, T.C., Chen, W.-H., Yu, K.L., Abdullah, R., 2018.
biomass gasification. Renewable and Sustainable Energy Reviews 38, 428–438. Torrefaction of microalgal biochar as potential coal fuel and application as bio-
Chang, G., Miao, P., Wang, H., Wang, L., Hu, X., Guo, Q., 2018. A synergistic effect adsorbent. Energy Conversion and Management 165, 152–162.
during the co-pyrolysis of Nannochloropsis sp. and palm kernel shell for aromatic Gao, N., Kamran, K., Quan, C., Williams, P.T., 2020. Thermochemical conversion of
hydrocarbon production. Energy Conversion and Management 173, 545–554. sewage sludge: A critical review. Progress in Energy and Combustion Science 79,
Chen, C., Ma, X., He, Y., 2012. Co-pyrolysis characteristics of microalgae Chlorella 100843.
vulgaris and coal through TGA. Bioresource Technology 117, 264–273. Gerardo, M.L., Oatley-Radcliffe, D.L., Lovitt, R.W., 2014. Integration of membrane
Chen, H., Wang, X., Wang, Q., 2020. Microalgal biofuels in China: The past, progress and technology in microalgae biorefineries. Journal of Membrane Science 464, 86–99.
prospects. GCB Bioenergy 12 (12), 1044–1065. Gong, X., Zhang, B., Zhang, Y., Huang, Y., Xu, M., 2014. Investigation on Pyrolysis of
Chen, H., Xie, Y., Chen, W., Xia, M., Li, K., Chen, Z., Chen, Y., Yang, H., 2019a. Low Lipid Microalgae Chlorella vulgaris and Dunaliella salina. Energy & Fuels 28
Investigation on co-pyrolysis of lignocellulosic biomass and amino acids using TG- (1), 95–103.
FTIR and Py-GC/MS. Energy Conversion and Management 196, 320–329. He, C., Tang, C., Li, C., Yuan, J., Tran, K.-Q., Bach, Q.-V., Qiu, R., Yang, Y., 2018. Wet
Chen, L., Yu, Z., Fang, S., Dai, M., Ma, X., 2018a. Co-pyrolysis kinetics and behaviors of torrefaction of biomass for high quality solid fuel production: A review. Renewable
kitchen waste and chlorella vulgaris using thermogravimetric analyzer and fixed bed and Sustainable Energy Reviews 91, 259–271.
reactor. Energy Conversion and Management 165, 45–52. He, Q., Guo, Q., Ding, L., Wei, J., Yu, G., 2019. Rapid co-pyrolysis of lignite and biomass
Chen, L., Yu, Z., Liang, J., Liao, Y., Ma, X., 2018b. Co-pyrolysis of chlorella vulgaris and blends: Analysis of synergy and gasification reactivity of residue char. Journal of
kitchen waste with different additives using TG-FTIR and Py-GC/MS. Energy Analytical and Applied Pyrolysis 143, 104688.
Conversion and Management 177, 582–591. Hu, Z., Ma, X., Li, L., 2016. The synergistic effect of co-pyrolysis of oil shale and
Chen, L., Yu, Z., Xu, H., Wan, K., Liao, Y., Ma, X., 2019b. Microwave-assisted co-pyrolysis microalgae to produce syngas. Journal of the Energy Institute 89 (3), 447–455.
of Chlorella vulgaris and wood sawdust using different additives. Bioresource Huang, Ming, Ma, Zhongqing, Zhou, Bingliang, Yang, Youyou, Chen, Dengyu, 2020.
Technology 273, 34–39. Enhancement of the production of bio-aromatics from renewable lignin by combined
Chen, R., Zhang, S., Yang, X., Li, G., Zhou, H., Li, Q., Zhang, Y., 2021a. Thermal approach of torrefaction deoxygenation pretreatment and shape selective catalytic
behaviour and kinetic study of co-pyrolysis of microalgae with different plastics. fast pyrolysis using metal modified zeolites. Bioresource Technology 301, 122754.
Waste Management 126, 331–339. https://doi.org/10.1016/j.biortech.2020.122754.
Chen, W.-H., Lin, B.-J., Lin, Y.-Y., Chu, Y.-S., Ubando, A.T., Show, P.L., Ong, H.C., Huang, Yanqin, Chen, Yupeng, Xie, Jianjun, Liu, Huacai, Yin, Xiuli, Wu, Chuangzhi,
Chang, J.-S., Ho, S.-H., Culaba, A.B., Pétrissans, A., Pétrissans, M., 2021b. Progress in 2016. Bio-oil production from hydrothermal liquefaction of high-protein high-ash
biomass torrefaction: Principles, applications and challenges. Progress in Energy and microalgae including wild Cyanobacteria sp. and cultivated Bacillariophyta sp. Fuel
Combustion Science 82, 100887. 183, 9–19.
Chen, W., Chen, Y., Yang, H., Xia, M., Li, K., Chen, X., Chen, H., 2017a. Co-pyrolysis of Istrate, I.-R., Iribarren, D., Gálvez-Martos, J.-L., Dufour, J., 2020. Review of life-cycle
lignocellulosic biomass and microalgae: Products characteristics and interaction environmental consequences of waste-to-energy solutions on the municipal solid
effect. Bioresource Technology 245, 860–868. waste management system. Resources, Conservation and Recycling 157, 104778.
Chen, W., Li, K., Xia, M., Yang, H., Chen, Y., Chen, X., Che, Q., Chen, H., 2018c. Catalytic Jonker, J.G.G., Faaij, A.P.C., 2013. Techno-economic assessment of micro-algae as
deoxygenation co-pyrolysis of bamboo wastes and microalgae with biochar catalyst. feedstock for renewable bio-energy production. Applied Energy 102, 461–475.
Energy 157, 472–482. Kasar, P., Sharma, D.K., Ahmaruzzaman, M., 2020. Thermal and catalytic decomposition
Chen, W., Lin, B.J., Huang, M.Y., Chang, J.S., 2015. Thermochemical conversion of of waste plastics and its co-processing with petroleum residue through pyrolysis
microalgal biomass into biofuels: a review. Bioresource Technology 184, 314–327. process. Journal of Cleaner Production 265, 121639.
Chen, W., Yang, H., Chen, Y., Li, K., Xia, M., Chen, H., 2018d. Influence of Biochar Katiyar, R., Arora, A., 2020. Health promoting functional lipids from microalgae pool: A
Addition on Nitrogen Transformation during Copyrolysis of Algae and review. Algal Research 46, 101800.
Lignocellulosic Biomass. Environmental Science & Technology 52 (16), 9514–9521. Klein, B.C., Bonomi, A., Maciel Filho, R., 2018. Integration of microalgae production
Chen, W., Yang, H., Chen, Y., Xia, M., Chen, X., Chen, H., 2017b. Transformation of with industrial biofuel facilities: A critical review. Renewable and Sustainable
Nitrogen and Evolution of N-Containing Species during Algae Pyrolysis. Energy Reviews 82, 1376–1392.
Environmental Science & Technology 51 (11), 6570–6579. Kumar, R., Strezov, V., Weldekidan, H., He, J., Singh, S., Kan, T., Dastjerdi, B., 2020.
Chiaramonti, D., Prussi, M., Buffi, M., Rizzo, A.M., Pari, L., 2017. Review and Lignocellulose biomass pyrolysis for bio-oil production: A review of biomass pre-
experimental study on pyrolysis and hydrothermal liquefaction of microalgae for treatment methods for production of drop-in fuels. Renewable and Sustainable
biofuel production. Applied Energy 185, 963–972. Energy Reviews 123, 109763.
Dai, G., Wang, S., Huang, S., Zou, Q., 2018a. Enhancement of aromatics production from Kuppens, T., Cornelissen, T., Carleer, R., Yperman, J., Schreurs, S., Jans, M., Thewys, T.,
catalytic pyrolysis of biomass over HZSM-5 modified by chemical liquid deposition. 2010. Economic assessment of flash co-pyrolysis of short rotation coppice and
Journal of Analytical and Applied Pyrolysis 134, 439–445. biopolymer waste streams. Journal of Environmental Management 91 (12),
Dai, L., Zhou, N., Li, H., Deng, W., Cheng, Y., Wang, Y., Liu, Y., Cobb, K., Lei, H., Chen, P., 2736–2747.
Ruan, R., 2020. Recent advances in improving lignocellulosic biomass-based bio-oil Lam, S.S., Wan Mahari, W.A., Ok, Y.S., Peng, W., Chong, C.T., Ma, N.L., Chase, H.A.,
production. Journal of Analytical and Applied Pyrolysis 104845. Liew, Z., Yusup, S., Kwon, E.E., Tsang, D.C.W., 2019. Microwave vacuum pyrolysis
Dai, M., Xu, H., Yu, Z., Fang, S., Chen, L., Gu, W., Ma, X., 2018b. Microwave-assisted fast of waste plastic and used cooking oil for simultaneous waste reduction and
co-pyrolysis behaviors and products between microalgae and polyvinyl chloride. sustainable energy conversion: Recovery of cleaner liquid fuel and techno-economic
Applied Thermal Engineering 136, 9–15. analysis. Renewable and Sustainable Energy Reviews 115, 109359.
Dai, M., Yu, Z., Fang, S., Ma, X., 2019. Behaviors, product characteristics and kinetics of Lee, Jechan, Kwon, Eilhann E., Park, Young-Kwon, 2020a. Recent advances in the
catalytic co-pyrolysis spirulina and oil shale. Energy Conversion and Management catalytic pyrolysis of microalgae. Catalysis Today. 355, 263–271.
192, 1–10. Lee, D.-J., Chang, J.-S., Lai, J.-Y., 2015. Microalgae–microbial fuel cell: A mini review.
Dasan, Y.K., Lam, M.K., Yusup, S., Lim, J.W., Lee, K.T., 2019. Life cycle evaluation of Bioresource Technology 198, 891–895.
microalgae biofuels production: Effect of cultivation system on energy, carbon Lee, D.-J., Lu, J.-S., Chang, J.-S., 2020b. Pyrolysis synergy of municipal solid waste
emission and cost balance analysis. Science of The Total Environment 688, 112–128. (MSW): A review. Bioresource Technology 123912.
del Río, P.G., Gomes-Dias, J.S., Rocha, C.M.R., Romaní, A., Garrote, G., Domingues, L., Levasseur, W., Perré, P., Pozzobon, V., 2020. A review of high value-added molecules
2020. Recent trends on seaweed fractionation for liquid biofuels production. production by microalgae in light of the classification. Biotechnology Advances
Bioresource Technology 299. 107545.

12
G. Su et al. Bioresource Technology 344 (2022) 126096

Li, K., Zhang, L., Zhu, L., Zhu, X., 2017. Comparative study on pyrolysis of lignocellulosic pretreatment methods on the exergetic aspects of lignocellulosic biofuels. Energy
and algal biomass using pyrolysis-gas chromatography/mass spectrometry. Conversion and Management 212, 112792.
Bioresource Technology 234, 48–52. Stamenković, O.S., Siliveru, K., Veljković, V.B., Banković-Ilić, I.B., Tasić, M.B.,
Liu, G., Wright, M.M., Zhao, Q., Brown, R.C., Wang, K., Xue, Y., 2016. Catalytic pyrolysis Ciampitti, I.A., Đalović, I.G., Mitrović, P.M., Sikora, V.Š., Prasad, P.V.V., 2020.
of amino acids: Comparison of aliphatic amino acid and cyclic amino acid. Energy Production of biofuels from sorghum. Renewable and Sustainable Energy Reviews
Conversion and Management 112, 220–225. 124, 109769.
Liu, R., Sarker, M., Rahman, M.M., Li, C., Chai, M., Nishu, Cotillon, R., Scott, N.R. 2020a. State, R.N., Volceanov, A., Muley, P., Boldor, D., 2019. A review of catalysts used in
Multi-scale complexities of solid acid catalysts in the catalytic fast pyrolysis of microwave assisted pyrolysis and gasification. Bioresource Technology 277,
biomass for bio-oil production – A review. Progress in Energy and Combustion 179–194.
Science, 80, 100852. Su, G., Ong, H.C., Ibrahim, S., Fattah, I.M.R., Mofijur, M., Chong, C.T., 2021. Valorisation
Liu, Y., Zhai, Y., Li, S., Liu, X., Liu, X., Wang, B., Qiu, Z., Li, C., 2020b. Production of bio- of medical waste through pyrolysis for a cleaner environment: Progress and
oil with low oxygen and nitrogen contents by combined hydrothermal pretreatment challenges. Environmental Pollution 116934.
and pyrolysis of sewage sludge. Energy 203, 117829. Su, Y., Song, K., Zhang, P., Su, Y., Cheng, J., Chen, X., 2017. Progress of microalgae
Lu, J.-S., Chang, Y., Poon, C.-S., Lee, D.-J., 2020. Slow pyrolysis of municipal solid waste biofuel’s commercialization. Renewable and Sustainable Energy Reviews 74,
(MSW): A review. Bioresource Technology 123615. 402–411.
Lv, X., Liu, H., Huang, Y., Yao, J., Yuan, H., Yin, X., Wu, C., 2021. Synergistic effects on Sun, C., Li, C., Tan, H., Zhang, Y., 2019a. Synergistic effects of wood fiber and polylactic
co-pyrolysis of low-temperature hydrothermally pretreated high-protein microalgae acid during co-pyrolysis using TG-FTIR-MS and Py-GC/MS. Energy Conversion and
and polypropylene. Energy Conversion and Management 229, 113772. Management 202, 112212.
Mei, Deqing, Guo, Dongmei, Wang, Cheng, Dai, Pengfei, Du, Jiayi, Wang, Junfeng, 2020. Sun, J., Xiong, X., Wang, M., Du, H., Li, J., Zhou, D., 2019b. Microalgae biodiesel
Evaluation of esterified pyrolysis bio-oil as a diesel alternative. Journal of the Energy production in China: A preliminary economic analysis. Renewable and Sustainable
Institute 93 (4), 1382–1389. Energy Reviews 104.
Menegazzo, M.L., Fonseca, G.G., 2019. Biomass recovery and lipid extraction processes Tang, F., Yu, Z., Li, Y., Chen, L., Ma, X., 2020a. Catalytic co-pyrolysis behaviors, product
for microalgae biofuels production: A review. Renewable and Sustainable Energy characteristics and kinetics of rural solid waste and chlorella vulgaris. Bioresource
Reviews 107, 87–107. Technology 299, 122636.
Meng, Haiyu, Wang, Shuzhong, Chen, Lin, Wu, Zhiqiang, Zhao, Jun, 2016. Study on Tang, Z., Chen, W., Chen, Y., Yang, H., Chen, H., 2019. Co-pyrolysis of microalgae and
product distributions and char morphology during rapid co-pyrolysis of platanus plastic: Characteristics and interaction effects. Bioresource Technology 274,
wood and lignite in a drop tube fixed-bed reactor. Bioresource Technology 209, 145–152.
273–281. Tang, Z., Chen, W., Hu, J., Li, S., Chen, Y., Yang, H., Chen, H., 2020b. Co-pyrolysis of
Meyer, Pimphan A., Snowden-Swan, Lesley J., Jones, Susanne B., Rappé, Kenneth G., microalgae with low-density polyethylene (LDPE) for deoxygenation and
Hartley, Damon S., 2020. The effect of feedstock composition on fast pyrolysis and denitrification. Bioresource Technology 311, 123502.
upgrading to transportation fuels: Techno-economic analysis and greenhouse gas life Tripathi, M., Sahu, J.N., Ganesan, P., 2016. Effect of process parameters on production of
cycle analysis. Fuel 259, 116218. https://doi.org/10.1016/j.fuel.2019.116218. biochar from biomass waste through pyrolysis: A review. Renewable and Sustainable
Mihalcik, D.J., Mullen, C.A., Boateng, A.A., 2011. Screening acidic zeolites for catalytic Energy Reviews 55, 467–481.
fast pyrolysis of biomass and its components. Journal of Analytical and Applied Uçar, S., Karagöz, S., Yanik, J., Saglam, M., Yuksel, M., 2005. Copyrolysis of scrap tires
Pyrolysis 92 (1), 224–232. with waste lubricant oil. Fuel Processing Technology 87 (1), 53–58.
Moens, L., Evans, R.J., Looker, M.J., Nimlos, M.R., 2004. A comparison of the Maillard Uzar, U., 2020. Political economy of renewable energy: Does institutional quality make a
reactivity of proline to other amino acids using pyrolysis-molecular beam mass difference in renewable energy consumption? Renewable Energy 155, 591–603.
spectrometry. Fuel 83 (11), 1433–1443. van Schalkwyk, D.L., Mandegari, M., Farzad, S., Görgens, J.F., 2020. Techno-economic
Mohajerani, A., Burnett, L., Smith, J.V., Markovski, S., Rodwell, G., Rahman, M.T., and environmental analysis of bio-oil production from forest residues via non-
Kurmus, H., Mirzababaei, M., Arulrajah, A., Horpibulsuk, S., Maghool, F., 2020. catalytic and catalytic pyrolysis processes. Energy Conversion and Management 213,
Recycling waste rubber tyres in construction materials and associated environmental 112815.
considerations: A review. Resources, Conservation and Recycling 155, 104679. Vanthoor-Koopmans, M., Wijffels, R.H., Barbosa, M.J., Eppink, M.H.M., 2013.
Nagappan, S., Devendran, S., Tsai, P.-C., Dinakaran, S., Dahms, H.-U., Ponnusamy, V.K., Biorefinery of microalgae for food and fuel. Bioresource Technology 135, 142–149.
2019. Passive cell disruption lipid extraction methods of microalgae for biofuel Varsha, S.S.V., Vuppaladadiyam, A.K., Shehzad, F., Ghaedi, H., Murugavelh, S.,
production – A review. Fuel 252, 699–709. Dong, W., Antunes, E., 2021. Co-pyrolysis of microalgae and municipal solid waste:
Nyoni, B., Duma, S., Bolo, L., Shabangu, S., Hlangothi, S.P., 2020. Co-pyrolysis of South A thermogravimetric study to discern synergy during co-pyrolysis process. Journal of
African bituminous coal and Scenedesmus microalgae: Kinetics and synergistic the Energy Institute 94, 29–38.
effects study. International Journal of Coal Science & Technology 7 (4), 807–815. Wang, S., Shang, H., Abomohra, A.E.-F., Wang, Q., 2019a. One-step conversion of
Ong, H.C., Chen, W.-H., Farooq, A., Gan, Y.Y., Lee, K.T., Ashokkumar, V., 2019. Catalytic microalgae to alcohols and esters through co-pyrolysis with biodiesel-derived
thermochemical conversion of biomass for biofuel production: A comprehensive glycerol. Energy Conversion and Management 198, 111792.
review. Renewable and Sustainable Energy Reviews 113, 109266. Wang, X., Ma, D., Jin, Q., Deng, S., Stančin, H., Tan, H., Mikulčić, H., 2019b. Synergistic
Ong, H.C., Chen, W.-H., Singh, Y., Gan, Y.Y., Chen, C.-Y., Show, P.L., 2020. A state-of- effects of biomass and polyurethane co-pyrolysis on the yield, reactivity, and heating
the-art review on thermochemical conversion of biomass for biofuel production: A value of biochar at high temperatures. Fuel Processing Technology 194, 106127.
TG-FTIR approach. Energy Conversion and Management 209, 112634. Wang, X., Zhao, B., Yang, X., 2016. Co-pyrolysis of microalgae and sewage sludge:
Ozyuguran, A., Akturk, A., Yaman, S., 2018. Optimal use of condensed parameters of Biocrude assessment and char yield prediction. Energy Conversion and Management
ultimate analysis to predict the calorific value of biomass. Fuel 214, 640–646. 117, 326–334.
Peng, X., Ma, X., Lin, Y., Guo, Z., Hu, S., Ning, X., Cao, Y., Zhang, Y., 2015. Co-pyrolysis Wu, Tu, J., Cai, Y., Wu, Z., Li, Z. 2020. Biofuel production from pyrolysis of waste
between microalgae and textile dyeing sludge by TG–FTIR: Kinetics and products. cooking oil fried sludge in a fixed bed. Journal of Material Cycles and Waste
Energy Conversion and Management 100, 391–402. Management, 22(4), 1163-1175.
Qi, P., Chang, G., Wang, H., Zhang, X., Guo, Q., 2018. Production of aromatic Wu, X., Wu, Y., Wu, K., Chen, Y., Hu, H., Yang, M., 2015. Study on pyrolytic kinetics and
hydrocarbons by catalytic co-pyrolysis of microalgae and polypropylene using behavior: The co-pyrolysis of microalgae and polypropylene. Bioresource
HZSM-5. Journal of Analytical and Applied Pyrolysis 136, 178–185. Technology 192, 522–528.
Resasco, D.E., Crossley, S.P., 2015. Implementation of concepts derived from model Wu, Zhiqiang, Yang, Wangcai, Li, Yaowu, Yang, Bolun, 2018a. Co-pyrolysis behavior of
compound studies in the separation and conversion of bio-oil to fuel. Catalysis Today microalgae biomass and low-quality coal: Products distributions, char-surface
257, 185–199. morphology, and synergistic effects. Bioresource Technology 255, 238–245.
Rezvani, S., Moheimani, N.R., Bahri, P.A., 2016. Techno-economic assessment of CO2 Wu, Zhiqiang, Yang, Wangcai, Yang, Bolun, 2018b. Thermal characteristics and surface
bio-fixation using microalgae in connection with three different state-of-the-art morphology of char during co-pyrolysis of low-rank coal blended with microalgal
power plants. Computers & Chemical Engineering 84, 290–301. biomass: Effects of Nannochloropsis and Chlorella. Bioresource Technology 249,
Sait, H.H., Salema, A.A., 2015. Microwave dielectric characterization of Saudi Arabian 501–509.
date palm biomass during pyrolysis and at industrial frequencies. Fuel 161, 239–247. Xiang, Zhongping, Liang, Jianghui, Morgan, Hervan Marion, Liu, Yuanyuan,
Sajjadi, B., Chen, W.-Y., Raman, A.A.A., Ibrahim, S., 2018. Microalgae lipid and biomass Mao, Hanping, Bu, Quan, 2018. Thermal behavior and kinetic study for co-pyrolysis
for biofuel production: A comprehensive review on lipid enhancement strategies and of lignocellulosic biomass with polyethylene over Cobalt modified ZSM-5 catalyst by
their effects on fatty acid composition. Renewable and Sustainable Energy Reviews thermogravimetric analysis. Bioresource Technology 247, 804–811.
97, 200–232. Xie, Q., Addy, M., Liu, S., Zhang, B., Cheng, Y., Wan, Y., Li, Y., Liu, Y., Lin, X., Chen, P.,
Sankaran, R., Parra Cruz, R.A., Pakalapati, H., Show, P.L., Ling, T.C., Chen, W.-H., Ruan, R., 2015. Fast microwave-assisted catalytic co-pyrolysis of microalgae and
Tao, Y., 2020. Recent advances in the pretreatment of microalgal and lignocellulosic scum for bio-oil production. Fuel 160, 577–582.
biomass: A comprehensive review. Bioresource Technology 298, 122476. Xin, C., Addy, M.M., Zhao, J., Cheng, Y., Cheng, S., Mu, D., Liu, Y., Ding, R., Chen, P.,
Sattar, Hamed, Muzaffar, Imran, Munir, Shahid, 2020. Thermal and kinetic study of rice Ruan, R., 2016. Comprehensive techno-economic analysis of wastewater-based algal
husk, corn cobs, peanut crust and Khushab coal under inert (N2) and oxidative (dry biofuel production: A case study. Bioresource Technology 211, 584–593.
air) atmospheres. Renewable Energy 149, 794–805. Yang, C., Li, R., Zhang, B., Qiu, Q., Wang, B., Yang, H., Ding, Y., Wang, C., 2019.
Sharara, M.A., Holeman, N., Sadaka, S.S., Costello, T.A., 2014. Pyrolysis kinetics of algal Pyrolysis of microalgae: A critical review. Fuel Processing Technology 186, 53–72.
consortia grown using swine manure wastewater. Bioresource Technology 169, Yang, W.Y., Zeng, Y., Luo, J., Tong, D., Qing, R.W., Fan, Y., hu, C. 2011. Production of
658–666. bio-oil by direct and catalytic pyrolysis of Nannochloropsis sp. Ranliao Huaxue
Soltanian, S., Aghbashlo, M., Almasi, F., Hosseinzadeh-Bandbafha, H., Nizami, A.-S., Xuebao/Journal of Fuel Chemistry and Technology, 39, 664-669.
Ok, Y.S., Lam, S.S., Tabatabaei, M., 2020. A critical review of the effects of

13
G. Su et al. Bioresource Technology 344 (2022) 126096

Yao, Z., Ma, X., Wu, Z., Yao, T., 2017. TGA–FTIR analysis of co-pyrolysis characteristics Zhang, Bo, Zhong, Zhaoping, Ding, Kuan, Song, Zuwei, 2015. Production of aromatic
of hydrochar and paper sludge. Journal of Analytical and Applied Pyrolysis 123, hydrocarbons from catalytic co-pyrolysis of biomass and high density polyethylene:
40–48. Analytical Py–GC/MS study. Fuel 139, 622–628.
Yu, D., Hui, H., Li, S., 2019. Two-step catalytic co-pyrolysis of walnut shell and LDPE for Zhang, J., Wang, K., Nolte, M.W., Choi, Y.S., Brown, R.C., Shanks, B.H., 2016. Catalytic
aromatic-rich oil. Energy Conversion and Management 198, 111816. Deoxygenation of Bio-Oil Model Compounds over Acid-Base Bifunctional Catalysts.
Yu, S., Cao, X., Wu, S., Chen, Q., Li, L., Li, H., 2020. Effective pyrolysis of waste cooking ACS Catalysis 6 (4), 2608–2621.
oils into hydrocarbon rich biofuel on novel mesoporous catalyst with acid and alkali Zhang, X., Lei, H., Zhu, L., Qian, M., Yadavalli, G., Wu, J., Chen, S., 2017. From plastics
coexisting. Industrial Crops and Products 150, 112362. to jet fuel range alkanes via combined catalytic conversions. Fuel 188, 28–38.
Yuan, T., Tahmasebi, A., Yu, J., 2015. Comparative study on pyrolysis of lignocellulosic Zhao, Y., Cao, H., Yao, C., Li, R., Wu, Y., 2020. Synergistic effects on cellulose and lignite
and algal biomass using a thermogravimetric and a fixed-bed reactor. Bioresource co-pyrolysis and co-liquefaction. Bioresource Technology 299, 122627.
Technology 175, 333–341. Zhu, X., Zhang, Y., Ding, H., Huang, L., Zhu, X., 2018. Comprehensive study on pyrolysis
and co-pyrolysis of walnut shell and bio-oil distillation residue. Energy Conversion
and Management 168, 178–187.

14

You might also like