Sie sind auf Seite 1von 162

Accepted Manuscript

Biomass pyrolysis: A review of the process development and


challenges from initial researches up to the commercialisation stage

Xun Hu , Mortaza Gholizadeh

PII: S2095-4956(18)30901-X
DOI: https://doi.org/10.1016/j.jechem.2019.01.024
Reference: JECHEM 767

To appear in: Journal of Energy Chemistry

Received date: 20 September 2018


Revised date: 7 January 2019
Accepted date: 30 January 2019

Please cite this article as: Xun Hu , Mortaza Gholizadeh , Biomass pyrolysis: A review of the process
development and challenges from initial researches up to the commercialisation stage, Journal of
Energy Chemistry (2019), doi: https://doi.org/10.1016/j.jechem.2019.01.024

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and
all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Highlights

 This study reviewed the progress in pyrolysis of biomass.

 Effects of process design and essential parameters on pyrolysis were


discussed.

T
IP
 Progress in reactor design, pretreatment of biomass, and catalytic pyrolysis

CR
were also discussed.


US
The key issues and fundamental challenges in pyrolysis of biomass were
analyzed.
AN
 Future research directions in pyrolysis of biomass were proposed.
M
ED
PT
CE
AC

1
ACCEPTED MANUSCRIPT

Review

Biomass pyrolysis: A review of the process development and challenges from

initial researches up to the commercialisation stage

Xun Hua, Mortaza Gholizadehb,


a
School of Material Science and Engineering, University of Jinan, Jinan 250022, Shandong, China

T
b
Faculty of Chemical and Petroleum Engineering, University of Tabriz, Tabriz, Iran

IP
CR
Abstract

US
Lignocellulosic biomass can be convert to a condensable liquid named bio-oil, a

solid product named as char and a mixture of gaseous products comprising CO2,
AN
CO, H2, CH4, etc. In recent years, much effort has been made on the investigation of

conversion of biomass through pyrolysis. However, commercialisation of the


M

biomass pyrolysis technology is still challenging due to various issues such as the

deleterious properties of bio-oil including the low heating value and the high
ED

instability at elevated temperatures. To overcome such issues, many processes,


PT

reactors and catalysts have been developed for pyrolysis and catalytic pyrolysis of

biomass. A state to the art of pyrolysis or catalytic pyrolysis of biomass need to be


CE

summarised to have an overall evaluation of the technologies, in order to provide a

useful reference for the further development of pyrolysis technology. This study
AC

reviews the various pyrolysis process, especially focus on the effects of essential

parameters, the process design, the reactors and the catalysts on the pyrolysis


Corresponding author. Tel: (+) 984133392990; Fax: (+) 984133338497; E-mail address:
m.gholizadeh@tabrizu.ac.ir (M. Gholizadeh).

2
ACCEPTED MANUSCRIPT

process. In addition, progress in commercialisation of pyrolysis technology was also

reviewed and the remaining issues in the process of commercialisation were

discussed.

Keywords: Biomass; Pyrolysis; Commercialisation; Current plants; Review

T
Xun Hu biography

IP
CR
US
Dr Hu obtained his PhD degree from the Chinese Academy of Sciences in 2010.
AN
From March 2010 to December 2012, he held a Postdoctoral Research Fellow

position in Curtin University (Australia) under the supervision of Professor Chun-Zhu


M

Li. From January 2013 to October 2016, he has held a Curtin Research Fellow
ED

position in Curtin University (Australia). Since October 2016, he works in University

of Jinan (China) as a full professor. His major research interest includes the
PT

development of various catalysts for hydrogenation of biomass-derived organics, for


CE

acid-catalyzed reactions and steam reforming reactions as well as the conversion of

biomass into fuels, chemicals and carbon materials, and the application of the
AC

functional carbon materials. From 2006 to 2018, he has published around 120

papers and has an h-index of 29.

3
ACCEPTED MANUSCRIPT

Mortaza Gholizadeh biography

T
IP
Dr. Gholizadeh was born in Ahar, Iran in 1978. He received his BSc (University of

CR
Tehran) in 2002 and MSc (Sharif University of Technology) in 2003. After that, he

worked in the refinery field for several years. Later he got PhD (Curtin University of

US
Technology) in 2015 in chemical engineering and then continued his study as a

postdoctoral researcher in Curtin University of Technology for one year. Recently he


AN
joint to University of Tabriz as an a/professor and continuing his research. So far, he

has published over 40 scientific papers. He has the h-index of 10. His main interest
M

is biomass pyrolysis and upgrading the bio-oil produced to biofuel by


ED

hydrodeoxygenation process.
PT
CE

1. Introduction

At the beginning of the 20th century, crude petroleum fuels covered only 4% of
AC

the world’s energy demand. However, nowadays, petroleum fuels are the most

important energy source and covers about 40% of the world’s energy demand. It

also produces 96% of the transportation fuels [1]. Nevertheless, petroleum fuels are

non-renewable and the reserves of fossil fuel are depleting fast. In addition, the use

4
ACCEPTED MANUSCRIPT

of petroleum fuels influences environment by generating huge amounts of net carbon

dioxide emission and other pollutants such as NOx and SOx. Therefore, there is an

urgent need to find renewable and environmentally benign feedstocks for sustainable

supply of fuels and energy.

Biomass, one of the potential feedstocks, meets such requirement. Biomass itself

T
is carbon-neutral and its use produces much less SOx and NOx due to the much

IP
lower content of nitrogen and sulphur in biomass than in coal or petroleum oil.

CR
Biomass can be converted to gaseous fuels via gasification or liquid fuels via

pyrolysis or hydrothermal liquefaction. Pyrolysis can be operated at atmospheric

US
pressure. Hence, the conversion of biomass into a liquid fuel via pyrolysis has

attracted great interest. Pyrolysis of biomass is categorised to slow and fast


AN
pyrolysis, according to the heating rate during pyrolysis. Fast pyrolysis is currently

the preferred route because of the fast reaction rate and the relatively higher yields
M

of bio-oil.
ED

Many researches have been done on fast pyrolysis including the studying of

mechanism of pyrolysis, reaction processes and design of reactors and also the
PT

development of catalysts for catalytic pyrolysis [2–4]. Many authors studied the
CE

reaction pathways of the component of biomass (cellulose, hemicellulose and lignin)

during pyrolysis [2–4]. Collard et al. proposed the mechanisms to explain the
AC

evolutions of some typical components of bio-oil. Three main mechanisms are

considered for pyrolysis of biomass including char formation, depolymerisation and

fragmentation. Secondary reactions such as cracking or recombination can also

occur. However, few researchers worked on the mechanism of real biomass

pyrolysis because of its complicate reaction network. Some researches on

5
ACCEPTED MANUSCRIPT

investigation of the kinetics for biomass pyrolysis were performed by using

thermogravimetric analysis (TGA), which was summarised by Wang et al. [2]. There

is a need for a comprehensive study to summarise all the proposed mechanism and

kinetics in model component of biomass or pyrolysis of biomass.

In addition to the reaction mechanism and reaction kinetics, understanding the

T
correlation of the product distribution with the process parameters during biomass

IP
pyrolysis is also a central interest in this research area. There are three major

CR
products from biomass pyrolysis, which includes char, bio-oil and pyrolytic gas. Char

is a solid product, the residual in pyrolysis of biomass, with low volatility and high

US
carbon content. Bio-oil is the primary product of interest, which is an organic mixture

of alcohols, ketones, aldehydes, phenols, ethers, esters, sugars, furans, alkenes,


AN
nitrogen and oxygen compounds. The pyrolytic gas is mainly the result of cracking

and decomposition of big molecules that forms from the initial stages of pyrolysis. It
M

consists of carbon dioxide, carbon monoxide, hydrogen, low carbon number


ED

hydrocarbons, nitrogen oxide, sulphur oxide and etc. The yield for char, bio-oil and

gas can vary significantly under the different process conditions. Typically, the yields
PT

of bio-oil, bio-char and the gaseous products are 50–70, 13–25 and 12–15 wt%,
CE

respectively [3].

A number of researches have been performed on investigating the compositions


AC

of char, bio-oil and pyrolytic gases in the pyrolysis of biomass or the model

components of biomass. Wang et al. summarised the product distributions in

pyrolysis of cellulose, hemicellulose and lignin [2]. Mohan et al. reviewed the bio-oil

composition produced from different groups of biomasses [3]. A number of

parameters can influence the pyrolysis products including their yields and

6
ACCEPTED MANUSCRIPT

compositions. These incudes the types of biomass and the conditions for the

pretreatment (physical, chemical and biological treatments), pyrolysis temperature,

heating rate, carrier gas type and etc. Nevertheless, how do the pyrolysis process

and the reaction parameters affect the formation of the organics in bio-oil, the

properties of bio-char and composition of pyrolytic gas needs to be further

T
elucidated, reviewed and summarised.

IP
Reactor is the heart of pyrolysis process. Several reactor types were used in the

CR
pyrolysis of biomass including fluidised bed, spouted fluidised bed, transported bed,

rotating cone, vortex centrifuge, augur or screw, radiative-convective entrained flow,

US
microwave, moving bed and fixed bed, ceramic ball downflow and vacuum ones [4].

Different reactors affect the yields, compositions or properties of char, bio-oil and
AN
pyrolytic gas in different ways. In addition to the above reactors, some new type of

reactors such as grinding pyrolyser was also developed and reached the stage of
M

demonstration. The different reactors had very different configurations and can
ED

substantially affect the process for pyrolysis. The effects of configuration of the

reactors on distribution of the pyrolysis products need to be understood. The


PT

correlation of the essential parameters of the reactors with especially the


CE

composition of bio-oil needs to be established.

Bio-oil produced from pyrolysis has some undesirable properties such as high
AC

corrosiveness, thermal instability, high oxygen content and correspondingly low

heating value. This makes it difficult for the direct use of bio-oil as an engine fuel. To

overcome these issues, bio-oil has to be upgraded via such as hydrodeoxygenation

of bio-oil. However, hydrodeoxygenation of bio-oil features with high pressure, which

induces the high cost of the process and other issues such as safety and catalyst

7
ACCEPTED MANUSCRIPT

deactivation, preventing this process of being commercialised [5–7].

Catalytic pyrolysis is another method developed to deoxygenate the bio-oil in situ

during the pyrolysis. In this process, the catalyst was directly added inside the

reactor to convert the heavy species inside the bio-oil to lighter ones [8–12]. Some

catalysts such as zeolite base ones were also tested or developed for catalytic

T
pyrolysis of biomass. Nevertheless, the overall performances of these catalysts in

IP
these separate studies need to be summarised and compared, in order to provide

CR
useful references for the further development of effective catalysts for the efficient

pyrolysis of biomass.

US
So far, some review papers in the biomass pyrolysis have been published,

focusing on specific aspects of the process. For example, Wang et al. and Collard et
AN
al. studied the mechanism for pyrolysis of biomass [2,13]. K Vamvuka reviewed the

properties of pyrolysis products [14]. Mohan et al. studied the chemical point of view
M

on the fast pyrolysis of lignocellulose feeds [3]. Bridgwater summarised the use of
ED

different reactors in the pyrolysis [4]. Collard et al. reviewed the model compounds

pyrolysis [13]. Czernik et al. studied the chemistry of the catalytic pyrolysis [12].
PT

Several other similar papers were also summarised specific aspect of the pyrolysis.
CE

In this review, a broad state-of-the-art multiple aspects of the pyrolysis is discussed.

The recent important results and trends in both fast and catalytic pyrolysis are
AC

summarised, focusing particularly on the mechanism of reactions, analysis of

products, effect of different parameters and different reactors and catalyst effects.

The optimum conditions were also summarised. Furthermore, the status of

commercialisation of biomass pyrolysis technology was analysed and the challenges

in the large scale pyrolysis were discussed.

8
ACCEPTED MANUSCRIPT

2. Products of biomass pyrolysis

Biomass pyrolysis is a complex process, determined by composition of biomass

feedstock, reaction parameters, reaction pathways of the main components and etc.

[13–20]. There are three major products from the biomass pyrolysis, which are bio-oil,

T
char and pyrolytic gas [20–25]. All these products are discussed in the following

IP
section.

CR
2.1. Bio-oil

US
Bio-oil as a product of interest from biomass pyrolysis is a dark brown organic

liquid. It is mainly composed of oxygenated compounds, which lead to the high


AN
thermal instability, and the low heating value, making it unusable as an engine fuel

[26]. Bio-oil is formed by simultaneous fragmentation and depolymerisation of


M

cellulose, hemicellulose and lignin in fast pyrolysis of biomass. Rapid heating of


ED

biomass and also fast quenching of vapour produced in fast pyrolysis results in the

production of bio-oil. On dry biomass base, the bio-oil yield from fast pyrolysis is 50–
PT

70 wt%, depending on reaction conditions [27].


CE

Bio-oil generally contains a high content of water and hundreds of organic

components such as acids, alcohols, ketones, furans, phenols, ethers, esters,


AC

sugars, aldehydes, alkenes, nitrogen and also oxygen compounds [28]. It also has

many reactive molecules and oligomeric species having molecular weight larger than

5000, which makes bio-oil unstable, even at room temperatures. Bio-oil also can be

considered as a multiphase microemulsion, in which the oligomers form aerosols

[29]. This makes it unstable, which is called aging. Aging leads to the formation of

9
ACCEPTED MANUSCRIPT

more water, higher viscosity and phase separation [29,30]. Therefore, bio-oil has to

be upgraded before the use as engine fuel. In Table 1, the typical properties of the

bio-oil produced from wood pyrolysis and heavy fuel oil are compared [3].

The produced bio-oil yield from different biomasses is normally in the range of

50–75 wt%. Normally, higher cellulose content enhances the yields of bio-oil, while

T
higher lignin content results in the formation of less bio-oil [26].

IP
The properties of bio-oil are dependent on a lot of parameters such as heating

CR
rate in the reactor, residence time inside the reactor, biomass particle size,

temperature and the type of biomass is used. They will be discussed later in details.

US
Osamaa et al. compared the different methods used for the analysis of bio-oil from

wood pyrolysis in different labs and the standard methods for pyrolysis of bio-oil
AN
were concluded [31]. In Table 2, the conclusion of their review is presented. The

water content in petroleum fuels should be regulated because it can form a separate
M

phase, resulting in corrosion, emulsion and some other problems in the burners. For
ED

the petroleum fuels, water can be easily separated by physical methods such as

centrifugation. However, in bio-oil, water exists as a micro emulsion and cannot be


PT

removed via physical methods. The water content of the bio-oil can be measured by
CE

Karl Fischer volumetric titration, according to ASTM Standard E 203 [31]. Bio-oil

typically contains some solids (<0.5 wt%) including condensed carbon residual
AC

material, sand and metals. The residual carbon can be measured according to

ASTM D4530 or ASTM D189, respectively [31]. CHN analysis of bio-oil can be

performed by ASTM Standard D 5291. It is worthwhile to note that using the above

mentioned method will result the carbon and hydrogen content with a fairly good

accuracy. However, the accuracy of nitrogen depends on several parameters such

10
ACCEPTED MANUSCRIPT

as its content, the detection limit of the equipment. For sulphur content determination,

ASTM Standard D5453-09 or EN ISO 20846 is recommended. So far, for chlorine

determination, no accurate method was found. Oxygen is typically calculated by

difference in this standard [31]. The specific gravity is used in calculating

weight/volume relationships. ASTM Standard D 4052 (EN ISO 3675) is normally

T
used to measure the density of the petroleum fuels at 15 °C, using a digital density

IP
meter. The same method is suggested to be used for the bio-oil density

CR
measurement. From the literature, it is found that ASTM Standard D 4052 had an

acceptable accuracy in the bio-oil density prediction [31]. The viscosity of the bio-oil

US
is another important specification of the bio-oil influencing the pumping condition.

From literature, the maximum kinematic viscosity of bio-oil was measured 20 cSt at
AN
40 °C. Depending on a lot of parameters such as temperature and water content, the

bio-oil viscosity can change significantly. Kinematic viscosity of bio-oil is measured


M

according to ASTM Standard D445 (EN ISO 3104) [31]. The pour point of a fuel is
ED

an important factor in pumping of that. It is typically shows the lowest temperature at

which the fuel can be pumped. For the fossil fuels, upper limit for pumpability is ∼600
PT

cSt. The same as petroleum fuels, the pour point of bio-oil can be determined
CE

according to ASTM Standard D 97 (EN ISO 3016). From literature, it is determined

that the pour point of bio-oil is typically below −20 °C [31]. The flash point of the bio-
AC

oil is measured by ASTM Standard D93, using a Pensky−Martens closed-cup tester.

However, this method is not a convenient method for the bio-oil. This is because the

bio-oil is made of light and heavy components, which can have a wide range of flash

point such as 40 up to 110 °C. As a result, determining a point for flashing of bio-oil

is complicated [31]. In general, for the most of the bio-oil properties, the petroleum

11
ACCEPTED MANUSCRIPT

fuel standards can be used. However, for some of the bio-oil properties, new

regulations and standards should be prepared.

2.2. Char

Char (also called biochar) is the solid product of biomass pyrolysis process,

T
which is a carbonaceous residue. Depending on biomass composition and pyrolysis

IP
process condition, char can have different chemical and physical properties. For

CR
instance, its carbon content could range from 53 wt% to 96 wt%. In addition, the

yield and heating value of char also varied in a wide range (30–90 wt% and 20–36

US
MJ/kg, respectfully) [32,33]. For instance, the char yields and elemental analysis

including surface area and micropore volume for eucalyptus are reported in Table 3
AN
[3].

Char can be supplied as a soil amendment including carbon sequestration, soil


M

fertility improvement, pollution remediation and agricultural by-product/waste


ED

recycling. It can also be used in catalytic utilisation, energy storage and

environmental protection and a sustainable platform carbon material for other high-
PT

value applications. Some researchers used it as a sorbent for the removal of


CE

pollutants in water. In addition, char also could be used as a sorbent for the removal

of pollutants in flue gas, such as SO2 and NOX [32].


AC

Char physical properties could also vary, depending on a lot of parameters. For

instance, the surface area undergoes some appreciable changes with temperature

(0.1–3.2 m2/g). In biomass pyrolysis process, below 400 °C, due to incomplete

removal of volatiles from biomass, surface area of the char products did not change

much. At 400–500 °C, the surface area of char starts increasing and the increase

12
ACCEPTED MANUSCRIPT

continues up to 900 °C. This increase could be up to 100–500 m2/g. The increase of

the surface area versus reaction temperature is because of the formation of

microporous structure [34]. In Table 4, the effect of pyrolysis temperature on some of

the physical properties of char produced from eucalyptus is shown [3].

Chemical specifications of char change substantially versus biomass type and

T
process conditions. For instance, functional groups present in char could change by

IP
the temperature of pyrolysis. The H/C and O/C for biomass is normally about 1.4–

CR
1.8 and 0.55–0.75, respectively. This shows that the biomass has low aromaticity

and has high aliphatic content. Pyrolysis leads to a significant decrease in the H/C

US
and O/C atomic ratios. The changes are more obvious with the increase of pyrolysis

temperature. Below 500 °C, the major reactions taking place during pyrolysis are
AN
mainly dehydration, decarboxylation and decarbonylation. However, above 500 °C,

dehydrogenation is the major reaction. The different reactions can affect the
M

functionality and the properties of char. The results from various researches show
ED

that biomass has the very complex transmittance bands such as hydroxyl, carbonyl,

ether group. From 185 to 200 °C, the functional group in biomass was not changed.
PT

However, at 200 to 500 °C, the dehydration of the intermediates from pyrolysis of
CE

biomass occurs. Simultaneously some aliphatic structures disappeared at 400–

550 °C. This was possibly because of the breaking of the weak bonds between the
AC

C–H of the alkyl groups with CH4 and C2-hydrocarbons formed. On the other hand, at

temperature higher than 300 °C, carboxyl and carbonyl group originating from

carbohydrates produced. In addition, at the temperature above 600 °C, the

abundance of aromatic rings having six or more fused benzene rings in the char

increased, indicating the aromatisation process [35,36]. Zhang et al. studied the

13
ACCEPTED MANUSCRIPT

effect of pyrolysis temperature on properties of the char produced from straw. In

Table 5, the carbon contents relating to different functional groups at different

temperatures are summarised [37].

2.3. Pyrolytic gas

T
The main gases produced in the pyrolysis of biomass include carbon dioxide,

IP
carbon monoxide, hydrogen, methane, ethane, ethylene, propane, sulphur oxides,

CR
nitrogen oxides, and ammonia. CO and CO2 are mainly originated from the

decomposition and also reforming of carboxyl and carbonyl groups. Light

US
hydrocarbons stemmed from the decomposition of weak methoxy and methylene

bonds. Hydrogen is the result of the reforming and decomposition of C–H groups and
AN
aromatics [38–41].

Reaction temperature has a significant influence on pyrolytic gas yield and


M

composition. As the pyrolysis temperature increases, thermal degradation and


ED

devolatilisation of biomass enhances. Simultaneously, the produced volatiles

undergo a series of secondary reactions such as decarboxylation, decarbonylation,


PT

dehydrogenetaion, deoxygenation and cracking to form pyrolytic gas [42]. Moisture


CE

content is another parameter influencing the pyrolytic gas production. High moisture

content can enhance the extraction of water-soluble components from the gaseous
AC

phase, hence causing a significant decrease in the yield of pyrolytic gas [43].

Particle size of biomass is another parameter influencing the yield and composition

of pyrolytic gas. Smaller particle size of biomass feedstock favours the cracking of

the components or the reactive intermediates, producing more H 2 and CO while less

CO2 [44].

14
ACCEPTED MANUSCRIPT

Vigouroux performed pyrolysis of straw, straw pellet, olive waste, birch and

bagasse at different temperatures. They observed that higher temperature favoured

the cracking of the reactive intermediates and consequently hydrogen yield

increased. In addition, the increase of temperature enhanced CO yield while that of

CO2 decreased. For instance, enhancing the temperature from 800 up to 1000 °C

T
during the pyrolysis of straw resulted in the reduced content of the hydrocarbon

IP
gases (CH4, C2H6, C2H4, C2H2) in the pyrolytic gas from 13 vol% to 4 vol%. However,

CR
the hydrogen content increased from 34 vol% to 44 vol%. Moreover, CO content

increased from 37 vol% to 47 vol% while CO2 reduced from 24 vol% to 5 vol%. The

US
particle size of biomass also significantly affects formation of the gases. With the

increase of the particle size of olive waste from 0.6 to 0.9 mm, the amount of
AN
hydrocarbon gas increased from 13 vol% to 18 vol%, while the hydrogen content

reduced from 31 vol% to 20 vol%. This was because the particle size influenced the
M

heating rate of the reactor. They concluded that the increase of particle size of
ED

biomass declined the heating rate and consequently increased the char yield. In

addition, smaller particle size favoured the cracking of hydrocarbons with an


PT

increase of hydrogen yield due to the longer resident time of volatiles in the reactor
CE

resulting higher amount of cracking reactions [44].


AC

3. Mechanism of pyrolysis

Depending on the heating rate and residence time of feedstock in reactor,

pyrolysis can be categorised to three main types including slow, fast and flash

pyrolysis [45,46]. Slow pyrolysis is mainly used for charcoal production. Its process

temperature could vary in the range of 300–700 °C while the biomass particle size of

15
ACCEPTED MANUSCRIPT

5–50 mm was used [47]. The heating rates for fast pyrolysis process is different from

slow one (heating rate > 10–200 °C/s, residence time = 0.5–10 s). In flash pyrolysis,

the heating of 103–104 °C/s with the biomass residence time less than 0.5 second is

used [48,49].

Moderate temperature (350–500 °C) and low residence time favours the

T
conversion of biomass into liquid products while low residence time and low

IP
temperature mainly converts the biomass into charcoal. In Table 6, the conditions for

CR
the liquid and charcoal production process are summarised [4].

A large number of reactions take place in the pyrolysis, including dehydration,

US
depolymerisation, isomerisation, aromatisation, decarboxylation and charring [13–15].
AN
The reactions could be categorised as below [13].

- Primary reactions: The primary reactions include char formation,


M

depolymerisation and fragmentation [16]. (a) Charring: Char is an aromatic polycyclic

carbon resulted from combination or condensation of benzene rings during pyrolysis


ED

[17]. (b) Depolymerisation: involves the cracking of the bond linkages between the

monomers. During pyrolysis, depolymerisation is a dominate reaction routes, which


PT

result in the production of volatiles and the gases [18,19]. (c) Fragmentation: The
CE

covalent linkage bonds of the unit monomers and polymer are fragmented. The

components with short chain and some incondensable gases are produced in the
AC

fragmentation of biomass [20,21].

- Secondary reactions: The primary compounds formed during the primary

reactions might not be stable and they can still undergo the secondary reactions

such as the cracking or recombination reactions. In addition, produced char can

catalyse the secondary reactions. From cracking of the primary molecules, lighter

16
ACCEPTED MANUSCRIPT

compounds will be produced while the recombination will result in the formation of

heavier compounds or deposit on surface of char [21,22].

- Reactions from different process conditions: Different process conditions can

substantially impact the reaction network. For instance, lower heating rate

(<10 °C/min) results in the production of more char, while fast heating (>100 °C/s)

T
produces more volatile compounds. In addition, the reaction temperature has a

IP
significant effect on the products distribution. At temperature 250–500 °C

CR
depolymerisation reactions rate is very high. From 450 up to 550 °C, the maximum

yield of bio-oil is achieved while above 550 °C, more fragmentation reactions happen,

US
forming more gaseous products [23–25]. Therefore, the process conditions in the

pyrolysis of biomass can significantly influence the reactions and the mechanism of
AN
the pyrolysis products formation, as discussed below. From all the studied have

been performed so far, it can be concluded that the secondary reactions are the
M

controlling reactions, which can build up the products with specific properties.
ED

3.1. Mechanism of the conversion of the components of biomass to the organic


PT

volatiles
CE

Three main components can be found in biomass: cellulose, hemicellulose and

lignin (Table 7). A small percentage of some resins and minerals are also present in
AC

the biomass components [50]. Pyrolysis is a complex process including a number of

reactions, which results in the formation of three main products including char, bio-oil

and gas. Each of the products is composed of some components stemming from

cellulose, hemicellulose and lignin. The experimental parameters such as

temperature, heating rate can affect the reaction pathway of the compounds [17]. As

17
ACCEPTED MANUSCRIPT

a result, studying the reaction mechanism for the formation of the different

compounds is desirable.

3.1.1. Cellulose conversion

Cellulose is composed of a linear homo polysaccharide of cellobiose monomers

T
and it includes two β-glucopyranose units connected by glycosidic bonds (Fig. 1).

IP
Depending on the source, the degree of polymerisation of the cellulose can reach

CR
more than 5000 units. In addition, cellulose is composed of crystalline (highly

ordered) and amorphous (randomly distributed) phases. Cellulose fibers provide the

US
strength of wood and its content in dry wood is 40–50 wt% [13].

The main products from pyrolysis of cellulose are acids, alcohols, anhydrosugars,
AN
char, gases and water. Initial kinetic model for cellulose pyrolysis was developed by

Bradbury et al. The kinetic model for each reaction was shown below [51]:
M

Cellulose → Active cellulose


ED

Activated cellulose → Char + Gas


PT

Activated cellulose → Tar → Gas

Yan-Fen et al. corrected Bradbury’s model and presented a modified model for
CE

the pyrolysis of cellulose (Fig. 2) [52]. Antal et al. used TGA to evaluate the

Bradbury’s model. For this aim, they used different biomasses and found that with a
AC

proper pretreatment of biomass, Bradbury’s model and proposed kinetic results are

similar to practical data [53]. Cellulose pyrolysis has also been studied in a laminar

entrained flow reactor by Brown et al. [54]. Different models’ results proposed in the

literature were compared by the data obtained in their experiment by them. They

18
ACCEPTED MANUSCRIPT

concluded that the Diebold model has the best predicts for general pyrolysis process.

However, it does not accurately predict char and secondary product formation from

the pyrolysis from high heating rate experiments conducted in Brown’s study. They

also concluded that Reynolds-Burnham rate predicted accurately the general

pyrolysis temperature regime but did not fit the data trend as well as the Diebold

T
model. In addition, the Antal-Varhegyi rate and classic (also known as Broido-

IP
Shafizadeh) model, which were principally derived from low heating rate experiments,

CR
predict that pyrolysis will occur at a much faster rate than has been observed [54].

Kinetics of cellulose pyrolysis after a pressurised heat treatment process was studied

US
by Wu et al. They pretreated cellulose by hot pressure treatment (300–350 °C and

10–15 MPa). The employed pretreatment resulted in the decomposition of the


AN
feedstock material in a wide temperature range. They also used distributed activation

energy model with Gaussian distribution for the pyrolysis kinetics [55]. Lin et al. used
M

both TGA and pyrolysis reactor experiments to study the cellulose pyrolysis. They
ED

conclude that the first step in the cellulose pyrolysis is the depolymerisation of

cellulose to form levoglucosan. Later levoglucosan could undergo dehydration and


PT

isomerisation reactions to form other anhydrosugars including levoglucosenone,


CE

1,4:3,6-dianhydro-β-D-glucopyranose and 1,6-anhydro-β-D-glucofuranose. The

anhydrosugars could react and form furans. Carbon monoxide and carbon dioxide
AC

were also produced from decarbonylation and decarboxylation reactions. In addition,

char is formed from polymerisation of the pyrolysis products. Two different reaction

models were proposed. The first model was the first-order kinetic model concluded

from TGA analysis. The second model combined the first-order kinetic model with an

energy balance in the pyroreactor [56]. Patwardhan at al. investigated the cellulose

19
ACCEPTED MANUSCRIPT

pyrolysis to clarify the primary and secondary reactions’ mechanisms. For this aim,

they used a micro reactor and also a bench scale fluidised bed reactor. Comparison

of the results from the two reactor systems revealed that in the fluidised bed reactor,

oligomerization of levoglucosan and decomposition of primary products such as 5-

hydroxymethyl furfural, anhydro xylopyranose and 2-furaldehyde were the major

T
secondary reactions [57]. Pyrolysis of cellulose at low temperature (300–350 °C) in a

IP
bench scale reactor was also studied by Wooten et al. By using Nuclear magnetic

CR
resonance spectroscopy (NMR) for analysis of the pyrolysis products they offered a

new model for cellulose pyrolysis (Fig. 3) [58]. Shen et al. proposed a mechanism for

US
the conversion of the cellulose in a fluidised bed pyrolyser (Fig. 4). Their model was

based on the comparison between levoglucosan and the other main products’
AN
reactions versus the possible routes for primary cracking of the cellulose molecules

and the secondary reactions of fragments [59].


M

From the results reported and discussed above, it can be concluded that at low
ED

temperature (<350 °C), the dominant reaction is the dehydration reaction, resulting in

the formation of char, gases and water. At 273–450 °C, cellulose was converted to
PT

anhydro sugares such as levoglucosan [16]. From TGA data, very little weight
CE

change of cellulose was observed up to 300 °C. The main conversion took place

between 300 and 390 °C. At the temperature higher than 400 °C, the aromatic
AC

residue remained [60]. The main reactions occur during that pyrolysis can be

categorised as following [13]: (a) formation of the anhydrocellulose or active

cellulose at 150–300 °C. (b) Depolymerisation reaction at 300–390 °C. (c) Charring

reaction at 380–800°C. The decomposition mechanism is shown in Fig. 5 [53].

20
ACCEPTED MANUSCRIPT

3.1.2. Hemicellulose conversion

Cellulose micro-fibrils are bound together by hemicellulose. It is

heteropolysaccharides with different compositions inside depending on biomass type.

The basic units of hemicellulose are composed of glucose, galactose, manose,

xylose, arabinose and glucuronic acid. Its content in biomass is considerable. As an

T
example, for dry wood, hemicellulose usually accounts for 25–35 wt% [61–63]. A

IP
schematic of hemicellulose molecule is shown in Fig. 6. During the pyrolysis of

CR
hemicellulose, similar decomposition reactions of cellulose at the same temperature

occur. However, comparing to cellulose, in TGA analysis at 200–260 °C, more

US
volatiles, less tar and char were produced. In addition, less levoglucosan and more

acetic acid were produced in the pyrolysis of hemicellulose [3].


AN
The initial kinetic model reported in the literature for hemicellulose pyrolysis is a

staged kinetic model, which is shown in Fig. 7 [64]. Table 8 provides the information
M

of process conditions used by different research groups worked on hemicellulose


ED

pyrolysis considering global model. A multi stage kinetic model is proposed by Miller

et al., as shown in Fig. 8. In this model, the formation of active hemicellulose is the
PT

first step of hemicellulose decomposition followed by two parallel reactions leading to


CE

the formation of either tar or char and gases [65]. Ranzi et al. proposed another multi

stage global kinetic model for hemicellulose pyrolysis (Fig. 9). In their model, two
AC

intermediate compounds were produced denoted as (C5H8O4)n, which was

subsequently converted to xylose, H2, H2O, CO, CO2, HCHO, CH3OH, C2H5OH and

char. Furthermore, G[CO2] and G[COH2] were also produced and later was

converted to CO2, CO, and H2. In this model, the possibility of any mass loss during

the pyrolysis of hemicellulose is considered. However, it could not predict the

21
ACCEPTED MANUSCRIPT

changes in product yields for different materials or even the same materials under

different pyrolysis conditions [66]. Xylan was also chosen as a model component to

predict the hemicellulose pyrolysis reactions. Chemical structure of xylan is shown in

Fig. 10 [67]. During the pyrolysis, the weight loss of xylan mainly happened at 220–

315 °C exothermically [68]. The pyrolysis of xylan produces acids, aldehydes, and

T
ketones, CO2, CO, CH4 and H2. Acetic acid and furfural were the most abundant

IP
products with the highest contents of almost 20 wt% [67].

CR
Concluded from the related literature, hemicellulose conversion mainly occurs in

the temperature range 200–350 °C. At temperatures higher than 350 °C,

US
rearrangement and char production are the main reactions [69]. The following

mechanism can be seen during the pyrolysis of hemicellulose [13]: (a) the linkages
AN
breakage and dehydration reactions at 150–270 °C; (b) depolymerisation reactions at
M

240–350 °C; (c) char formation at 350–800 °C. The chemical reactions of xylan

pyrolysis are shown in Fig. 11 [53]. In another research performed by Wang et al.,
ED

the mechanism of xylose pyrolysis was also studied. Their study is summarised in

Fig. 12 [67].
PT
CE

3.1.3. Lignin conversion

Lignin is the strengthening component of the biomass cell wall, which is mainly
AC

made of the polymer of hydroxyl and methoxy substituted propyl phenol units [61].

Depending on the type of biomass, the proportion of lignin monomers is different. In

soft wood it accounts for 23–33 wt% of the weight while in hard wood lignin content

is 16–25 wt% [26]. A schematic of lignin molecular is shown in Fig. 13. Through to

22
ACCEPTED MANUSCRIPT

TGA analysis, the main decomposition reactions of lignin occur at 200–450 °C. The

highest conversion is at 360–400 °C [70–72].

Lignin is composed of phenyl propane units, which are p-hydroxyphenyl (H),

guaiacyl (4-hydroxy-3-methoxyphenyl, G) and syringyl (3,5-dimethoxy-4-

hydroxyphenyl, S) [73]. Various kinds of linkage can be seen in lignin structure such

T
as ether linkages (C−O) and condensed (C−C) bonds, resulting in a heterogeneous

IP
chemical structure for lignin. Both TGA and bench scale reactors were used to study

CR
lignin decomposition during the pyrolysis. Through TGA analysis, three main peaks

for lignin pyrolysis were seen: the peak at 350 °C was for primary pyrolysis reactions

US
while at 400–450 °C methoxyl group-related reactions occurred. The last peak

appeared at 550–600 °C relating to gasification of catechols (1,2-dihydroxybenzenes)


AN
from secondary pyrolysis reactions [73].

From lignin pyrolysis, gas, solid and liquid products can be achieved. The
M

gaseous products include CO, CO2, H2, CH4, C2H4, C2H6, C3H6, C4H8, HCHO,
ED

CH3CHO. CO, H2 and CO2 are the most abundant component in the gas phase [74].

Ferdous et al. measured that ∼25 mol% of gas produced during the pyrolysis of
PT

lignin was H2 [75,76]. Solid product of lignin pyrolysis was studied by several
CE

researchers. Their studies indicated that char has an aromatic structure and 50% of

biomass energy stays in char [77]. Sharma et al. characterised the char produced
AC

from lignin pyrolysis by SEM, FTIR, and CPMAS 13C NMR. Their results showed

that the cleavages of aliphatic OH, carboxyl, and methoxyl groups have been

improved by enhancing the pyrolysis temperature [78]. In another research, Hosoya

et al. indicated that methoxyl groups is the most important functionalities influencing

the char formation during lignin pyrolysis [79]. Chu et al. proposed that the formation

23
ACCEPTED MANUSCRIPT

of char was due to polymerisation of small radical species and the elimination of side

functional groups including hydroxyl group [80]. Jimenez et al. indicated that

softwood lignin yielded guaiacyl derivatives, coniferaldehyde, and coniferyl alcohol

as the major products. The pyrolysis of hardwood lignin gave guaiacyl and syringyl

derivatives, syringaldehyde, coniferyl alcohol, and sinapyl alcohol. Pyrolysis of

T
bamboo lignin produced pvinylphenol as the major compound [81]. Jiang et al.

IP
analysed the liquid product of lignin pyrolysis in the range of 673–1073 K. Their

CR
results showed that the maximum yield of phenolic compounds was obtained at 873

K [82]. Greenwood et al. found that liquid product is mainly composed of guaiacol, 4-

US
methyl-guaiacol, vinylguaiacol, eugenol, vanillin, coniferylaldehyde, vinylguaiacol,

syringol, 3,5-dimethoxyacetophenone, 4-methyl 2,5-dimethoxy benzaldehyde, 4-allyl-


AN
dimethoxylphenol, syringaldehyde, 2,6-dimethoxyl-2-propylphenol, and
M

sinapaldehyde [83]. Jegers et al. could measure guaiacol, 4-methylguaiacol, 4-

ethylguaiacol, catechol, 4-methylcatechol, 4-ethylcatechol, phenol, cresol, and 4-


ED

ethylphenol in the liquid product of pyrolysis of lignin [84]. Lou et al. pyrolysed lignin

at different temperatures. Their results indicated that the content of methoxyl


PT

contained components was lower at elevated temperatures. However, the contents


CE

of non-methoxyl contained compounds increased with enhancing the temperature

[85]. Liu et al. revealed that lignin undergoes three main reactions during the
AC

pyrolysis. Initially, water evaporates followed by the formation of primary volatiles

and finally small molecular gases were produced [86].

Custodis et al. studied the mechanism of lignin fast pyrolysis. They used a

microreactor connected to the gas chromatograph with a mass spectrometer (py-

GC/MS) detects the final products. Diphenylether (DPE) and ortho-methoxyphenol

24
ACCEPTED MANUSCRIPT

(guaiacol) were selected as model compounds of lignin. Their results showed that

radical fragments such as hydroxycyclopentadienyl radical in guaiacol decomposition

are in an agreement with the Franck−Condon simulation. In addition, the degree of

substitution plays a dominant role in both the stabilisation of the intermediate radicals

[87]. The pyrolysis of four types of lignin including milled amur linden wood lignin

T
(MWL), enzymatic hydrolysis corn stover lignin (EHL), wheat straw alkali lignin (AL)

IP
and wheat straw sulphonate lignin (SL) was studied by Lin et al. They concluded that

CR
different lignins had different thermolysis behaviours. However, the lignins had

similar functional groups according to the FTIR analysis. In addition, syringyl,

US
guaiacyl and p-hydroxyphenylpropane structural units were broken down during

pyrolysis. The product distributions of the components inside the products from
AN
pyrolysis depended strongly on the lignin origin and isolation process. Phenols were

the major products from MWL, EHL and AL pyrolysis. However, SL produced a large
M

number of furan compounds and sulphur compounds [88]. An international study of


ED

fast pyrolysis of lignin was undertaken by Nowakowski et al. Thermogravimetric

analysis and a bench-scale test included bubbling fluidised-bed reactors were used.
PT

Based on the results, it was concluded that a concentrated lignin (estimated at about
CE

50% lignin and 50% cellulose) behaved like a typical biomass producing bio-oil.

However, pure lignin produced a much lower amount of bio-oil [89]. Brebu et al.
AC

studied thermal degradation of aspen lignin, lignosulfonate and cellolignin. They

found that the thermal behavior definitely depends on the lignin type. Aspen lignin

decomposed to acetic acid, methanol and methylacetate at 200–340 °C and also

phenol derivatives were formed at 250–450 °C. Cellolignin, a solid fraction resulted

during furfural manufacturing [90]. Fast pyrolysis of lignin by a micro-pyrolyser was

25
ACCEPTED MANUSCRIPT

investigated by Patwardhan et al. Their system had a residence time of 15–20 ms

declining the secondary reactions. Their results showed that monomeric compounds

are the primary pyrolysis products of lignin recombining after primary pyrolysis to

produce oligomeric compounds [91]. A two pathway model by Brebu et al. was

proposed by for lignin decomposition during the pyrolysis, which was shown in Fig.

T
14 [92].

IP
Concluding all in lignin pyrolysis, it is an amorphous three dimensional resin with

CR
a cross-linked structure. It decomposes mainly at temperature 280–500 °C. During

the pyrolysis, from the cleavage of ether and carbon-carbon bonds in lignin, phenols

US
yield. Furthermore, comparing to cellulose, more char from lignin is produced. The

TGA analysis of lignin revealed that its decomposing begins at 280 °C and continues
AN
up to 500 °C with the maximum rate at 350–450 °C [3]. The main reactions of lignin

pyrolysis are [13]: (a) Alkyl chains conversion including linkage breaking between
M

units at 150–420 °C; (b) Aromatic rings’ conversion and char formation reactions at
ED

380–800 °C. The mechanism of lignin pyrolysis is shown in Fig. 15 [74].


PT

3.2. Mechanism of biomass conversion to the products


CE

Biomass conversion during the pyrolysis consists of three main reactions (char

formation, depolymerisation and fragmentation) and secondary reactions. Lignin


AC

mainly results in the formation of char and also phenolics. Anhydro-saccharides and

furan stems from polysaccharides decomposition. The secondary reactions produce

gas and small components in the liquid and gas products. Collard et al. summarised

the three reactions mechanism, which is shown in Fig. 16 [13]. The authors

described three main reaction routes for the primary pyrolysis of biomass. In the first

26
ACCEPTED MANUSCRIPT

step, the solid residue with an aromatic polycyclic structure (char) formation was

explained. Via this pathway, benzene rings were formed. In addition, the combination

of these rings to produce a polycyclic structure was demonstrated. In addition, water

and the non-condensable gases were the by-products of the reactions resulting in

the formation of char. In the second route, the depolymerisation of the chains in the

T
molecules is shown. The depolymerisation reactions continue until a volatile

IP
compound is formed. The resulting molecules were mainly condensable at ambient

CR
temperature and were present in bio-oil. The third route is determining the breaking

of the linkages of covalent bonds. Non-condensable gases and also small chain

US
compounds could be formed in this step [13]. Greenhalf et al. studied the pyrolysis of

different biomass including straw, perennial grasses and hardwoods. They found that
AN
hardwoods need higher temperature for decomposition [93]. Qu et al. studied the

effect of biomass type on the composition of the products. Their study on peanut
M

straw and corn stalk pyrolysis indicated that peanut straw produced more
ED

carbohydrates while corn stalk resulted in more aldehydes due to more

hemicellulose content inside [94]. Butler et al. also studied the pyrolysis of spruce,
PT

salix, miscan-thus and wheat straw. They could find out that the lignin structure
CE

influenced the products from that. For instance, more guaiacol products from

guaiacyl units for spruce, syringol products from syringyl units for salix and lignin-
AC

derived phenols from p-hydroxyphenyl units for miscanthus were measured [95].

Biagini et al. could prove by simulation that the interaction of the main components of

biomass during the pyrolysis can be ignored and studying model compounds can

give an acceptable result for whole biomass pyrolysis [96]. Hosoya et al. proved that

higher content of lignin in the biomass accelerates the polymerisation and

27
ACCEPTED MANUSCRIPT

carbonisation reaction [97]. Liu et al. conclude that the presence of lignin decreased

the 2-furaldehyde content in the product [98]. The study of Wang et al. proved that

hemicellulose-cellulose interaction could increase the formation of 2,5-

diethoxytetrahydrofuran [99]. Zhang et al. proved that there is no significant

interaction between cellulose and hemicellulose during the pyrolysis while the

T
interaction of cellulose and lignin is considerable during the biomass pyrolysis [100].

IP
TGA analysis and bench scale reactors were used to reveal the biomass

CR
pyrolysis mechanism. However, according to the literature, the exact reaction

mechanisms of biomass pyrolysis are not clear completely because of not clear

US
understanding of the main components of biomass interaction during pyrolysis. One-

component mechanism was proposed by Shafizadeh et al. for wood pyrolysis. The
AN
summary of their mechanism is shown in Fig. 17 [101]. Branca et al. proposed a

multi-component mechanism for biomass pyrolysis shown in Fig. 18 [102].


M
ED

4. Different pretreatment processes to enhance the pyrolysis efficiency

To enhance the pyrolysis efficiency, the biomass needs to be pretreated. The


PT

pretreatment can increase the process efficiency through the destruction of


CE

lignocellulosic structure. Several typical methods for the pretreatment of biomass are

discussed in the following sections.


AC

4.1. Physical pretreatment

As biomass is not a good heat conductor, there is a temperature gradient inside it

influencing pyrolysis mechanism. The bigger biomass particle size, the more char is

produced in the pyrolysis. In addition, the reduction of biomass particle size

28
ACCEPTED MANUSCRIPT

increases the operation cost. Therefore, the selection of correct particle could have

an important impact on pyrolysis performance considering the cost [45].

4.1.1. The effects of particle size of biomass on pyrolysis

The effect of biomass particle size on the pyrolysis process and costs was

T
studied by several research groups. Their results indicated that bigger particle size

IP
results in more char and gas formation while the yield of bio-oil declines. In addition,

CR
particle size increase after a certain amount is not influencing the yield of the

products. More interestingly, the yield of water produced increases by the increase of

US
the particle size [103–105]. Bennadji et al. studied the effect of particle size on the

yields of products by using TGA and also a bench scale tubular reactor. Their results
AN
showed that particle size of biomass can have a strong effect on devolatilisation

timing and also influenced the yields of some species [105]. Rapeseed was
M

pyrolysed by Sensoz et al. in a heinze reactor at 500°C to monitor the particle size
ED

(0.224–1.8 mm) effect on the products. Their result indicated that the yields of char

and bio-oil produced are independent of particle size. The products’ composition
PT

showed a slight change by rapeseed particle size [106]. Lu et al. studied effects of
CE

particle shape and size on devolatilisation of biomass particle. Their study showed

that the particle size and shape influenced the reaction rate. In addition, the yields of
AC

the volatiles were reduced by decreasing the particle size. In addition, particle shape

influences the yields and composition of the products. For instance, spherical shape

biomass produced less volatiles than cylinder and plate shape biomasses [107]. Ali

et al. studied the effect of particle size of maize stalk on the yield of bio-oil. A

fluidised bed reactor with the temperature of 360–540 °C was used. The biomass

29
ACCEPTED MANUSCRIPT

particle size was 1–2 mm. It was observed that the particle size 1 mm at 490 °C

produced the bio-oil with maximum yield (42 wt%) [108]. Effect of particle size and

temperature on woody biomass fast pyrolysis at high temperature (1000–1400 °C)

was investigated by Septien et al. They used a drop tube furnace to perform the

pyrolysis experiments. Particle size was chosen in the range of 0.35–0.80 mm. Their

T
results showed that particle size did not have any effects on the products from

IP
pyrolysis [109]. Effects of particle size on the fast pyrolysis of oil mallee woody

CR
biomass were studied by Shen et al. Mallee wood particle size was 0.18–5.60 mm

and the reactor type was fluidised bed. The yield of bio-oil decreased because of

US
increasing the particle size from 0.3 to 1.5 mm. Further increases in biomass particle

size did not reduce the bio-oil yield. They concluded this as mainly due to the impact
AN
of particle size in the production of lignin derived compounds. It was also found that
M

the yields of light species in the bio-oil increased and those of heavy species inside

the bio-oil decreased with increasing biomass particle size. Changes in biomass cell
ED

structure during the reducing of particle size may also influence the yield and

composition of bio-oil [110]. Suriapparao et al. investigated the effects of biomass


PT

particle size on slow pyrolysis kinetics and fast pyrolysis product distribution. They
CE

used mixed wood sawdust with eight different particle sizes (26.5–925 μm) at

different heating rates of 0.5–100 °C min−1. They observed that the yield of phenolics
AC

and linear hydrocarbons decreased and the production of gases like CO and CO 2

increased when particle size increased during fast pyrolysis. High yield of aromatics

was measured with medium sized particles (362.5, 512.5 μm) [111].

30
ACCEPTED MANUSCRIPT

4.1.2. Densification

The density of the biomass can influence the pyrolysis products yields and also

compositions. In some researches, the palletisation of fractionated agricultural

straws and high pressure densification of wood residues to produce an economical

T
feedstock for pyrolysis reactor were studied. Their results confirmed that higher bio-

IP
oil yield can be achieved by compacting biomass [112–114].

CR
Common systems for biomass densification include pellet mill, cuber, briquette

press, screw extruder, tabletizer, and agglomerator. For pyrolysis process the pellet

US
mill, briquette press, and screw extruder are the most common ones.
AN
4.2. Chemical pretreatment

Biomass contains as the inorganic minerals, which can significantly influence the
M

pyrolysis process. Their effects could be decreased by chemical pretreatment.


ED

4.2.1. Acid and alkali pretreatment


PT

Carbonates, phosphates, sulphates and chlorides are the typical species found in
CE

the biomass, which influences the pyrolysis considerably. For instance, metals such

as sodium largely remained in char after pyrolysis and causes problem during the
AC

char combustion process. It can also increase the corrosiveness to the reactor

[65,66]. Cations also catalyse the char formation reactions and negatively affect bio-

oil formation. In addition, their presence in bio-oil accelerates the aging process of

bio-oil [118–120].

31
ACCEPTED MANUSCRIPT

The common method to remove the minerals from biomass before pyrolysis is

acid treatment. In this method, diluted acids such as sulphuric or chloric acid are

used [121]. The acid treatment could increase the yield of bio-oil from 19 to 27 wt%

[122]. Acid can also dissolve or remove hemicellulose, which is the source of

producing hydroxyl acids [122–124]. Mainly acid pretreatment is composed of two

T
types, including the one at high temperature (above 180 °C) for short duration (1–5

IP
min) and the one at low temperature (<120 °C) for long duration (30–90 min). Saha

CR
et al. concluded that due to the generation of high amount of inhibitory products such

as furfurals, 5-hydroxymethylfurfural, phenolic acids, and aldehydes, acid treatment

US
of biomass is problematic [125]. Effect of diluted sulphuric acid on pine wood

pretreatment was also studied by Wang et al. Their results showed that the
AN
concentrations of furfural and 5-hydroxymethylfurfural in the bio-oil from acid

pretreated pinewood were less than that from the untreated one [127]. The pyrolysis
M

of corncob pretreated by sulphuric acid was conducted by Wang et al. They also
ED

showed that acid pretreatment influenced liquid product distribution. For instance,

the yield of furfural reached as high as 77 wt% when the mass ratio of acid and
PT

corncob was 0.1:1. Levoglucosenone with the yield of 8.7 wt% was also produced
CE

with the acid to corncob ratio of 0.01:1. In addition, the high-value chemicals such as

4-vinyl guaiacol, 2,3-dihydrobenzofuran and 1,2-cyclopentanedione were produced


AC

when acid pretreatment was used [128]. Marzialetti et al. also investigated the effect

of the pretreatment with the different acids including trifluoroacetic acid, HCl, H2SO4,

HNO3, and H3PO4 on the pyrolysis of loblolly pine in a batch reactor. They concluded

that the biomass pretreated with trifluoroacetic acid produced the highest amount of

32
ACCEPTED MANUSCRIPT

soluble monosaccharides during pyrolysis of biomass under the pretreatment

condition of 150 °C and pH 1.65 [131].

Alkaline solutions such as NaOH with high concentration at low temperature were

also used to remove partial hemicellulose and lignin from biomass. During the

pretreatment, acidic compounds may react with the alkaline, resulting in the

T
formation of salt, which can catalyse the formation of char and consequently the

IP
yield of bio-oil decreases [132–134].

CR
Suruzzaman studied the two-step pretreatment of beech wood with peracetic acid

and NaOH, respectively [137]. The TGA results showed that thermal degradation

US
occurred during pyrolysis. In addition, majority of hemicellulose and lignin were

removed from the biomass during the two-step pretreatment of the beech wood
AN
firstly with peracetic acid and then NaOH. The char yield was also decreased for the

pretreated beech wood sample [137]. Wang et al. chose 0.5 wt% NaOH to pretreat
M

pinewood. Their results showed that bio-oil yield from NaOH-treated pinewood, was
ED

lower than from untreated pinewood. In addition, lignin content was reduced by

NaOH pretreatment [127]. In another research, Wang et al. used NaOH with 0.0002–
PT

0.2 mol/L to pretreat corncob. They selected a ratio of 0.08:1 for NaOH :corncob.
CE

Their result showed that the yield and composition of bio-oil significantly decreased

by NaOH pretreatment of corncob. They could conclude that the pretreatment by


AC

base could promote the cracking reactions during the pyrolysis, which resulted in

higher and lower yields of gas and bio-oil, respectively [128].

33
ACCEPTED MANUSCRIPT

4.2.2. Hydrothermal pretreatment

To increase the carbon and energy content of biomass, hydrothermal

pretreatment (also called wet torrefaction) of biomass is also employed. The

temperature in the range of 18–260 °C with the pressure up to 4.69 MPa was

T
typically used in this process. During hydrothermal process, the compounds such as

IP
acetic acid, pentoses, hexoses, furfural, phenolic compounds, sugars are produced.

CR
Almost whole hemicellulose decomposes while cellulose partially degrades during

the thermal treatment. Depolymerisation and carbonisation of hemicellulose, lignin,

US
and cellulose could occur during the torrefaction of biomass. The torrefaction

products includes solid, liquid (including water, organics, and lipids) and gases (CO 2,
AN
CO, and CH4). Typically, after torrefaction, the solid product contains 90% of the

initial energy content. The loss of the OH functional group during torrefaction makes
M

the biomass hydrophobic and consequently resistant towards chemical oxidation and
ED

microbial degradation [118,140–142]. In Fig. 19, a schematic of torrefaction

pretreatment of biomass for the production of bio-oil is shown [143].


PT

Different types including rotating drum, spiral reactor, herreshoff oven or multiple
CE

hearth furnace, moving bed, belt conveyor, microwave, fluidised bed etc. reactors

were used in biomass torrefaction. During the torrefaction, biomass is heated directly
AC

or indirectly by means of water vapour as a heating source. The process can be

controlled by means of temperature, rotating speed, length and angle of the drum

[144]. The advantages and disadvantages of the reactors are summarised in Table 9

[145].

34
ACCEPTED MANUSCRIPT

Zheng et al. studied the effect of temperature on the structure of torrefied

biomass. They chose the condition of 210, 230, 250, 270 and 300 °C with 30 min

residence time for torrefying the biomass. The torrefied biomass was used as a feed

of fluidesed bed pyrolysis at 450 °C. Their results showed that torrefaction can

improve significantly the bio-oil properties. In addition, by increasing the torrefaction

T
temperature, the –OH and C=O contents decreased. The char yield also declined by

IP
increasing temperature while the bio-oil yield consequently increased. The phenolics

CR
content in the bio-oil increased. However, the abundance of oxygen-containing

compounds such as acids, sugars and furans decreased, indicating that the

US
properties of bio-oil were improved by torrefying the biomass [146]. Meng et al.

studied the effect of biomass torrefaction on properties of the bio-oil produced. They
AN
concluded that by using torrefied biomass, the O/C ratio in bio-oil reduced [147].

Chen et al. demonstrated that the yield of char increased by enhancing the
M

torrefaction temperature while bio-oil yield decreased. Furthermore, oxygen content


ED

in the bio-oil also decreased [149]. In another research, Chen et al. studied the

torrefaction effect on pyrolysis behaviour of biomass and also the properties of the
PT

bio-oil produced by using TG-FTIR and Py-GC/MS. They demonstrated that the
CE

pyrolysis characteristics and kinetics properties of torrefied rice husk at 290 °C were

unique. In addition, fast pyrolysis of the torrefied rice husk produces a bio-oil with low
AC

moisture content and high heat value [151]. Torrefied biomass was used as a feed

for fast pyrolysis by Louwes et al. They proved that torrefaction could enhance the

efficiency of fast pyrolysis by decreasing the required energy for grinding biomass

particles and also it improves the properties of bio-oil. Their reactor was a down flow

reactor with 500 g/h feeding rate. They also selected wood as their reactor feed.

35
ACCEPTED MANUSCRIPT

Their results showed that the oxygen content of wood declined while its energy value

increased after the torrefaction. Furthermore, hardwood pellets torrefied at 265 °C

with a residence time of 45 min produced a bio-oil with the lowest oxygen content

and also its energy content increased from 19.1 MJ kg-1 to 23.1 MJ kg-1 [152].

Neupane et al. investigated the effect of torrefaction on biomass structure and bio-oil

T
produced from fast pyrolysis. For this aim, they used a Py-GC/MS. Temperatures of

IP
225, 250 and 275 °C and residence times of 5, 30 and 45 min were selected. The

CR
pyrolysis in the absence or presence of HZSM-5 catalyst was performed. They

proved that torrefaction caused deacetylation and decomposition of hemicellulose

US
and demethoxylation of lignin, resulting in an overall increase in aromaticity of

biomass. For non-catalytic pyrolysis, abundance of the phenolics in bio-oil increased


AN
with increasing the torrefaction severity while abundance of the furan compounds

decreased. In catalytic pyrolysis, they torrefied the biomass for 15 min at 225 and
M

250 °C. It increased the yield of aromatic hydrocarbon and also total carbon yield
ED

comparing to catalytic pyrolysis of the non-torrefied biomass [153]. The effects of

torrefaction on compositions of bio-oil and syngas from sawdust pellet pyrolysis by


PT

microwave heating were studied by Ren et al. They observed that the concentrations
CE

of phenols and sugars in bio-oil increased. However, the torrefaction reduced the

concentrations of guaiacols and furans. Torrefaction also reduced CO2 content while
AC

increased H2 and CH4 contents in syngas. This showed that the quality of syngas

was significantly improved by torrefaction process [154]. Ru et al. studied the effect

of torrefaction on physicochemical properties of biomass and the influence on

pyrolysis behaviour of biomass. They could observe that torrefaction reduced

hemicellulose content. It also increased the heating value of the biomass. The main

36
ACCEPTED MANUSCRIPT

reactions during torrefaction were dehydration, deacetylation, and cleavage of ether

linkages. They also found that the yields of acids and furans in bio-oil produced

decreased while more phenolics were produced [155].

4.2.3. Steam explosion

T
In this process, the fibres in biomass were opened up and its polymers become

IP
more accessible for subsequent processes, i.e. fermentation, hydrolysis or

CR
densification processes. Because of the tightly packed cellular structures (fibres) in

the biomass, the mechanical strength especially woody biomass is high. The rigid

US
structure of biomass is a major technical obstacle for most biorefinery processes

including pyrolysis. Steam explosion can extract the sugars and other useful
AN
compounds from biomass. It can also produce biomass pellets to increase the

calorific value. The effect of steam explosion on biomass composition and its effect
M

on pyrolysis were investigated by many researchers. In their studies, normally wood


ED

chips are fed from a bin through a screw loading valve and they are heated by using

steam at a temperature of about 285 °C and a pressure of 1–3.5 MPa for about 2
PT

min. This makes an explosive decompression of the biomass that results in a rupture
CE

of the rigid fibres structure. Different residence times and temperatures can result in

anything from small cracks in the wood structure, to total defibrillation of the wood
AC

fibres. Acetic acid can also be released during the steam explosion, and this result in

partial hydrolysis of the cell wall components. Steam explosion pretreatment does

not involve the use of chemicals, which is an advantage compared to other pre-

treatments. Therefore, steam explosion pre-treatment of biomass prior to

37
ACCEPTED MANUSCRIPT

pelletisation can increase heating value, bonding properties and hydrophobicity of

the wood [2,156,157].

Steam explosion and its combinatorial pretreatment refining technology of plant

biomass to bio-based products were reviewed by Chen et al. Their review suggested

that the combination of steam explosion and other pretreatments could significantly

T
influence the properties of biomass and modified the product distribution during the

IP
subsequent pyrolysis or hydrolysis [170].

CR
4.2.4. Ammonia fibre expansion

US
Ammonia fibre expansion is an important pretreatment method in which both

physical (high temperature and pressure) and chemical (ammonia) processes is


AN
used to achieve effective pretreatment. Anhydrous ammonia, high pressure (~2 MPa)

and temperature 60–120 °C were used for ammonia fibre expansion pretreatment of
M

biomass. A schematic of a common reactor used for ammonia fibre expansion is


ED

presented in Fig. 20 [176]. During the pretreatment, hemicellulose degraded into

oligomeric sugars while the depolymerisation of lignin occurred as well while


PT

cellulose recrystallised [2].


CE

4.3. Thermal pretreatment - drying effect


AC

The drying of biomass prior to pyrolysis is used for energy conservation. The

water in biomass requires a considerable amount of energy to vaporise during

pyrolysis. Decreasing the water content in the biomass will reduce the amount of

energy required to increase the feedstock temperature to the process temperature

with a higher ramping rate. Due to the short residence time in fast pyrolysis, drying is

38
ACCEPTED MANUSCRIPT

a critical stage, which normally occurs at the ambient temperatures or up to 150 °C.

On the other hand, water content is an important parameter due to its significant

effect on char formation [143,183]. Therefore, to reduce energy consumption and

also to produce less water in bio-oil, drying should be considered as an important

pretreatment in pyrolysis.

T
Some researchers investigated the effects of drying of biomass on pyrolysis

IP
products. Wang et al. studied the effect of microwave drying of biomass on biomass

CR
pyrolysis. Their study showed that the yields of char and bio-oil increased while the

yield of the gas product decreased by drying the biomass prior to the pyrolysis. In

US
addition, CO2 concentration in gas increased. In the char, lower organic compound

content was obtained [184]. Chen et al. also studied the effect of drying biomass on
AN
devolatilisation. They proved that moisture had no influence on the components and

structure of biomass [185].


M
ED

4.4. Biological pretreatment

Biological pretreatment is another environmental friendly method which uses rot


PT

fungi to decompose some of the biomass components to make it easier to process in


CE

pyrolysis [186]. Various types of rot fungi including brown, white and soft rot ones

can be used. They mainly degrade lignin and hemicellulose in the biomass. The
AC

main side products from this pretreatment are water and carbon dioxide. By this

method, the biomass becomes compact and more suitable for pyrolysis as a matter

of energy saving [187]. In addition, rot fungi declines the emission of SOX which is an

environment pollutant gas [89]. Different types of rot fungi can influence biomass in

39
ACCEPTED MANUSCRIPT

distinct ways. For instance, the white rot fungi have substantial lignin degrading

ability while the brown one promotes the degradation of cellulose [187].

Yang et al. used three types of white rot fungus to pretreat the corn stover. The

study of bio-treated biomass pyrolysis in TGA showed that bio-pretreatment could

decrease the pyrolysis temperature (1–35 °C). In addition, the emission of SOX

T
declined for 30%–45% [189]. Under the similar condition, Yu et al. employed ZSM-5

IP
zeolite catalyst for the pyrolysis, resulting in the production of 10 wt% higher content

CR
of valuable aromatics with 20 wt% reductions in the yield of coke formed [190].

Effects of biological pretreatment on pyrolysis behaviours of corn stalk by TG-FTIR

US
technique were studied by Wang et al. The results revealed remarkable influence of

biological pretreatment on the pyrolysis behaviour of corn stalk. They concluded that
AN
the changes of the chemical structure and components of corn stalk before and after

the pretreatment were the main reason of its different behaviour [194]. Pyrolysis-GC-
M

MS to assess the fungal pretreatment efficiency for wheat straw anaerobic digestion
ED

was investigated by Rouches et al. For this aim, white rot fungi Polyporus Brumalis

BRFM 985 and wheat straw were used. The results were monitored with Py-GC-MS.
PT

They observed that fungal pretreatment was required for anaerobic digestion of
CE

lignocellulosic biomass because of the increase in the amount of methane

production [200]. Yan et al. investigated pyrolysis characteristics and kinetics of


AC

lignin derived from enzymatic hydrolysis residue of bamboo pretreated with white rot

fungus. They concluded that the structural changes of lignin during pretreatment

could promote the formation of the guaiacyl-type derivatives during the fast pyrolysis.

Therefore, enzymatic hydrolysis of lignocellulose residue could impact pyrolysis

process [202].

40
ACCEPTED MANUSCRIPT

In Table 10, the summary of advantages and disadvantages of different physical

and chemical biomass pretreatment processes is presented.

5. Investigating the effects of reaction conditions to enhance the pyrolysis

efficiency

T
IP
5.1. Reaction atmosphere

CR
Pyrolysis of biomass is always carried out in an inert atmosphere. N 2, H2, CO,

CO2, CH4 and steam can be used as a carrier gas in the pyrolysis. The use of steam

US
has advantages. It can increase the yield of bio-oil through to decreasing the

secondary cracking reactions’ rates [204]. Therefore, carrier gas can change the
AN
mechanism of the pyrolysis reaction by influencing of the products’ yields, functional

groups in the products, energy content value of the bio-oil and etc. For instance, in
M

one research different carrier gases including N2, CO, CO2, CH4 and H2 were used
ED

for pyrolysis of biomass in a fluidised bed reactor. Their result indicated that CH4 and

CO produced the highest (58.7 wt%) and lowest (49.6 wt%) bio-oil yields,
PT

respectively [205].
CE

Zhang et al. investigated the pyrolysis of biomass in N2, CO2, CO, CH4 and H2

atmospheres. They concluded that the yields and compositions of the products
AC

depended on the type of carrier gas. The CO2 atmosphere resulted in less char than

in the other atmospheres. Less CO2 was also produced in CO2 atmosphere than in

N2 atmosphere. In addition, the CO2 atmosphere led to the highest yield of acetic

acid. It also produced more Ketones while phenolics content declined. The reason

for this could be either due to the reaction of CO2 with the active volatiles or with

41
ACCEPTED MANUSCRIPT

char [205]. Effects of adding CO2 to the atmosphere of biomass fast pyrolysis (N2)

were studied by Guizani et al. Their study showed that the addition of CO2 to the

pyrolysis atmosphere influenced on all the products yields and composition. Mixing

CO2 with nitrogen medium enhanced CO production. This was because of the

homogeneous and heterogeneous reactions of CO2 with the pyrolysis products. It

T
resulted in a lower char yield comparing the char yield with pure nitrogen [206].

IP
Kwon et al. also observed that the introducing of CO2 in the pyrolysis atmosphere

CR
enhanced gas yield while the oil yield reduced [207]. Atmospheric effects on

microwave pyrolysis of corn stover were investigated by Huang et al. They used

US
microwave power level of 500 W for 30 min in the pyrolysis process (N2 and CO2

atmosphere). Their results proved that under N2 atmosphere, the pyrolysis efficiency
AN
was higher than that under CO2 atmosphere. This was because of the higher heat

absorbability of CO2 molecules declining the heat transfer for pyrolysis. In addition, in
M

the gas product CO was the most abundant component under N 2 atmosphere while
ED

CO2 was the main component under CO2 atmosphere [208]. Hanaoka et al. studied

char properties produced under N2/CO2/O2 atmosphere. They observed that


PT

comparing to pure CO2 atmosphere, the mixed atmosphere resulted in a more


CE

developed surface area and a higher reactivity towards pure CO 2. However, similar

char yields were obtained in both atmospheres [209]. The effect of pyrolysis
AC

atmosphere on bio-oil yield and structure was investigated by Onal et al. Almond

shell was selected as a feedstock. They observed that bio-oil yield increased by

replacing the static atmosphere to nitrogen and steam. Moreover, co-feeding steam

during the pyrolysis increased aliphatics and reduced asphaltenes contents in the

bio-oil. In addition, using steam atmosphere enhanced H/C and heating value of bio-

42
ACCEPTED MANUSCRIPT

oils [210]. Zhang et al. studied the influence of reaction atmosphere (N2, CO, CO2

and H2) on microwave-induced fast pyrolysis of medicinal herb residue over ZSM-5

catalyst. Their study showed that CO atmosphere produced the lowest and highest

yields of bio-oil and water, respectively. The hydrocarbon contents in bio-oil under

the four reaction atmospheres indicated the following order: CO > CO2 > H2 ≈ N2.

T
However, the oxygen contents in bio-oil showed an opposite tendency. In addition,

IP
CO atmosphere resulted in the production of a more stable bio-oil. Furthermore, CO2

CR
atmosphere produced the highest yield of syngas (CO + H 2) which might be formed

via the reverse water gas shift reaction [211]. Borrego et al. pyrolysed wood under

US
N2 and CO2 atmospheres to study the properties of the char produced. They found

similar specific areas for char produced under N2 and CO2 atmospheres [212].
AN
Depending on the biomass type, reactor type and also process conditions, in the

same atmosphere different results for the yields and compositions of pyrolysis
M

products can be achieved. This is because of the possibility of the reacting of the
ED

reaction atmosphere with biomass under the process condition. From literature, it is

proved that except the inert gases, the others can react with biomass in pyrolysis
PT

process. For instance, CO2, can decay the residual solids comparing to an inert
CE

atmosphere such as N2 [212]. CO2 seems to have an affinity to react with

hydrogenated and oxygenated groups, leading to a more carbon rich char. In


AC

addition, the flowrate and consequently the amount of the non-inert gas can

influence the products’ yields and compositions [212]. From the literature, steam

atmosphere can increase the yield of organic oxygenates by preventing some of the

secondary reactions. CH4 atmosphere results in a high yield of bio-oil. Using CO as

a pyrolysis reaction atmosphere produces more stable bio-oil due to the less

43
ACCEPTED MANUSCRIPT

production of the methoxy-compounds. CO2 atmosphere produces a carbon rich

char. Finally, H2 atmosphere results in a fuel with a high heating value. Concluding

all above, it is crucial to add non-inert gases to the inert atmosphere of the pyrolysis

process. Depending on the aim of the biomass pyrolysis for producing bio-oil, char or

gas product, the atmosphere of the pyrolysis can be chosen. It seems that the

T
presence of the CO can be also very magnificent to produce a chemically stable bio-

IP
oil.

CR
5.2. Temperature

US
During the pyrolysis, decomposition of biomass proceeds via several

consecutive stages. Some internal rearrangement such as water elimination, bond


AN
breakage, appearance of free radicals, formation of carbonyl, carboxyl and

hydroperoxide groups takes place at 122 to 202 °C [213]. In the second stage (200–
M

600 °C), the pyrolysis products including char, bio-oil and gases start to form. The
ED

third stage (> 600 °C) is the decomposition of char at a very slow rate, and therefore,

carbon-rich residual solid forms [214]. The composition of the pyrolysis products also
PT

depends on temperature. At temperatures higher than 600 °C, the contents of the
CE

polar, aliphatic and aromatic compounds increase, resulting from the dehydration

and decarboxylation reactions. The yields of the different gas compounds change as
AC

well [40]. Amount of the water produced is also significantly affected by

temperature. At the temperature up to 360 °C, the production of water increased

significantly and while the further increase of temperature did not lead to the

increase of water production. However, from 580 °C, the water production starts to

decrease [215].

44
ACCEPTED MANUSCRIPT

Weerachanchai, et al. investigated the influence of temperature on the product

yields from the pyrolysis of palm kernel cake and cassava pulp residue. They

observed that by increasing the temperature, the yield of char declined [216]. Zanzi

et al. observed that an increase in carbon content with a decrease in hydrogen and

oxygen contents by enhancing the pyrolysis temperature [217]. Rahman et al.

T
studied the effect of temperature on characterization of pyrolysis products from oil

IP
palm fronds. They observed that the maximum yield of char was obtained at 300 °C

CR
while the maximum yield of bio-oil was obtained at 400 °C. Increase of temperature

resulted in more acidic bio-oil with a higher ash content [218]. Effect of temperature

US
on pyrolysis products from biomass was evaluated by Demirbas et al. They

concluded that the amount of char decreases with increasing the pyrolysis
AN
temperature. The highest bio-oil yield was obtained between 650 and 800 K. In
M

addition, heating value of the fuels produced was correlated with pyrolysis

temperature [219]. Ibrahim et al. studied the influence of reaction temperature on


ED

wheat straw pyrolysis. A maximum yield of bio-oil was achieved at 525–550 °C

[220].The effect of pyrolysis temperature on the products from acacia wood was
PT

evaluated by Noumi et al. Charcoal yield decreased with an increase in temperature


CE

of pyrolysis [221]. Rover et al. investigated the effect of pyrolysis temperature on

recovery of bio-oil from red oak. In their study, the maximum char yield (31.1 wt%)
AC

was achieved at 350 °C while the maximum bio-oil yield (66.7 wt%) was obtained at

400 °C. In addition, the highest non-condensable gas yield (26.3 wt%) was obtained

at 550 °C and also the maximum content of sugar from cellulose and hemicellulose

(13.5 wt%) was produced at 450 °C [222]. Effects of pyrolysis temperature and

residence time on physicochemical properties of different char types were studied by

45
ACCEPTED MANUSCRIPT

Sun et al. Pyrolysis temperatures including 300 °C, 400 °C, 500 °C and 600 °C were

selected. They observed that fixed carbon content, pH value and amount of basic

functional groups in char enhanced at the temperature from 300 °C to 600 °C while

the char yield, adsorption capacity and amount of acidic functional groups were

reduced correspondingly [223]. Effects of pyrolysis temperature on the

T
physicochemical properties of char derived from vermicompost and its potential use

IP
as an environmental amendment was studied by Yang et al. They concluded that the

CR
char yield decreased as the pyrolysis temperature increased [224]. Zhang et al.

studied the effects of pyrolysis temperature and heating time on char obtained from

US
the pyrolysis of straw and lignosulfonate. They observed that with the increase of

pyrolysis temperature, pH, ash content, carbon stability and total content of carbon
AN
were enhanced. However, char yield, volatile matter, and total content of hydrogen,

oxygen, nitrogen and sulphur declined. In addition, porosity and aromaticity of char
M

increased with the increase of pyrolysis temperature [37]. The study by Zhou et al.
ED

showed that the primary products of lignin pyrolysis were oligomers. Between 450

and 700 °C, the yield of the oligomers was very high. Mono phenols and other light
PT

compounds were obtained by secondary reactions inside the particle or in the vapour
CE

phase [225].
AC

5.3. Heating rate

The heating rate of biomass particles is the main parameter to differ slow

pyrolysis and fast pyrolysis. In slow pyrolysis, the heating rate range is 1–100 °C/min

while the heating rates higher than 1000 °C/min are required for fast pyrolysis

processes [3,4,226]. Depolymerisation of cellulose and hemicellulose could be

46
ACCEPTED MANUSCRIPT

promoted by increasing the heating rate. It also minimise the residence time of the

volatiles inside the reactor and also secondary reactions. In addition, the higher

heating rate promoted the cracking reactions and produced more volatiles (bio-oil)

while less char [227].

On the other hand, low temperatures and low heating rates will increase intra-

T
chain hydrogen bonds of cellulose functional groups, enhancing the probability of

IP
collision to produce a dehydration reaction [228].

CR
The composition and properties of bio-oil, char and gas compounds are also

influenced by heating rate. The higher heating rates decrease the water content in

US
the bio-oil due to the inhibition of secondary reactions such as volatile dehydration.

The CO and CO2 also increase at high heating rates. With high heating rates, a char
AN
with smaller pore volume is produced [229].

Weerachanchai et al. showed that products’ yields did not change significantly by
M

changing the heating rate in the range of 5–20 °C/min [216]. Putun et al. also proved
ED

that the change of heating rate in the range of 7–40 °C/min had no important effect

on the pyrolysis products’ yields [230]. On the other hand, Karaosmanoglu et al.
PT

observed that the heating rate higher than 30 °C/min resulted in higher yields of
CE

liquids and gas products and consequently lower char yield [231]. Angin studied the

effect of heating rate on the char properties produced from the pyrolysis of safflower
AC

seed press cake. Their study showed that altering heating rate did not influence the

heating value of the char produced [232]. Haykiri-Acma evaluated the effect of

heating rate on the pyrolysis yields of rapeseed. Their results showed that at the

lower heating rate, the maximum rates of mass losses were low. However, by

enhancing the heating rate, the maximum rates of mass losses increased. They

47
ACCEPTED MANUSCRIPT

concluded that this was because of mass or heat transfer resistance inside the

biomass particles at the low heating rate. At high heating rate, these restrictions

decreased and the conversion rate enhanced [233]. The study of Li et al. indicated

that at a low heating rate (<15 °C/min), the bio-oil mainly composed of molecules

with the weight range of 500–1000 while at higher heating rate (15 °C/min) it was at

T
the range of 200–500. Increasing the heating rate (>15 °C/min) did not change the

IP
weight of bio-oil anymore [234]. Effect of heating rate on the char produced from the

CR
pyrolysis of rapeseed was studied by Zhao et al. They showed that the heating rate

had a non-linear relation with char yield. With the increase of the heating rate from 1

US
to 5 °C/min, the char yield increased. However, from 5 to 20 °C/min, the char yield

declined and the optimum heating rate for char production was at ~5 °C/min [235].
AN

5.4. Vapour residence time


M

The products’ yields and property from biomass pyrolysis are significantly
ED

affected by the residence times of vapour and solid inside the reactor. Shorter

residence time of vapour favours bio-oil production and minimise cracking reactions
PT

[3,4]. On the other hand, higher residence time for biomass is needed to ensure
CE

complete decomposition [236].

Zanzi et al. studied the effect of residence time on product distributions from
AC

pyrolysis. Longer residence time of biomass inside the reactor may lead to the

occurrence of the secondary reactions, producing secondary products [217]. Aquino

et al. observed that by increasing the pyrolysis time during batch pyrolysis of cotton

gin trash, the char yield decreased slightly. In addition, the gas yield increases with

enhancing the pyrolysis time [237]. Gan et al. concluded that the increase of

48
ACCEPTED MANUSCRIPT

residence time did not enhance the gas yield significantly. In addition, they reported

a drastic decline of the char yield with increase in retention time [238]. The effect of

residence time on biomass pyrolysis was discussed by Cai et al. In their study, TGA

was used to study the pyrolysis kinetics characteristics of biomass at different

retention time. They proved that with the increase of the residence time, the cracking

T
reactions occurred more and the content of heavier species in the bio-oil declined.

IP
[239]. Fast pyrolysis behaviour of banagrass at different volatiles residence time in a

CR
fluidised bed reactor was investigated by Morgan et al. For this aim, four residence

times between ~1.0 and 10 s were selected and pyrolysis product distributions of

US
bio-oil, char and gas were determined under each condition. They observed that the

highest bio-oil yield was obtained at volatiles residence time of 1.4s (37 wt% yield on
AN
dry biomass base). At the same condition, the yields of char gas were 4 wt% and 8

wt%, respectively. In addition, longer residence times resulted in less bio-oil while
M

more char was produced and higher yield of gas and light volatiles were obtained
ED

[240]. Newalkar et al. investigated the effect of biomass residence time on the

pyrolysis product distributions in an entrained flow reactor. Heating rates of


PT

103−104 °C/s were achieved with the residence time ranging from 4 s to 28 s.
CE

Differences in the release of volatiles, evolution of char morphology and

carbonisation of the char skeleton were observed at different retention times. In


AC

addition, CO, CO2, H2 and CH4 were the major gas products evolved. Moreover,

C2−C4 hydrocarbons, oxygenates and benzene were the minor species produced. At

a longer residence time, higher content of polynuclear aromatic tars produced, which

was due to gas phase molecular weight growth reactions [241]. The study of Sun et

al. showed that enhancing the residence time at 300 °C reduced the char yield and

49
ACCEPTED MANUSCRIPT

increased the pH and iodine adsorption number of the char. In contrast, increasing

the residence time at 600 °C had little influence on the char yield and pH. However, it

decreased the iodine adsorption number of the char [223]. Zhao et al. observed that

the surface area and morphology of char were significantly influenced by residence

time. In addition, they concluded that residence time (10 to 100 min) did not

T
influence the char yield from rapeseed pyrolysis, indicating that even at a short

IP
residence time, maximum char yield was achievable [235].

CR
6. Reactors

US
Reactor is the major part of any pyrolysis process. The heating system is the

important section of the process facilities. To achieve this goal, depending on the
AN
method of heating system (Fig. 21), several reactors were established during the last

two decades. In Table 11, an overview of fast pyrolysis reactor characteristics for
M

bio-oil production is presented. The major reactor designs can be categorised as


ED

below [4,26].
PT

6.1. Bubbling fluidised beds


CE

Bubbling fluidised bed pyrolysers’ performance is consistent with a high liquid

yields of typically 70–75 wt% from wood on a dry feed basis. The biomass particle
AC

sizes should be small and less than 2–3 mm to achieve high biomass heating rates.

In fluidised bed reactors, the heating rate of particle is usually the limiting step.

Fluidising gas flow rate controls the residence time of solids and vapours. Normally

the residence time of char is higher than vapours. After pyrolysis, the char produced

should be separated from bio-oil as it acts as an effective vapour cracking catalyst.

50
ACCEPTED MANUSCRIPT

For this aim, usually an ejection and entrainment system is installed followed by

separation in one or more cyclones [3,4]. A common schematic of bubbling fluidised

bed reactor is shown in Fig. 22.

University of Waterloo in Canada started the initial research on bubbling fluid

reactor [242–244]. Dynamotive built a pilot plant with a capacity of 75 kg/h and 400

T
kg/h [245]. Wellman designed a 250 kg/h pilot plant in the UK [246]. Anhui University

IP
of Science and Technology constructed three pilot plants up to 600 kg/h capacity

CR
[247].

Bubbling fluidised bed reactor was used by Ali et al. to study the effect of

US
operation parameters on the pyrolysis product from maize stalk. Operation

conditions were: temperature range of 360–540 °C, feed particle size of 1–2 mm and
AN
carrier gas flow rate of 7.0–13.0 m3/h. They concluded that the maximum yield of

bio-oil was 42 wt% produced at 490 °C with the particle size of around 1.0 mm and
M

carrier gas flow rate of 11.0 m3/h [108]. Wood pyrolysis in a bubbling fluid reactor
ED

was simulated by Seok et al. For this aim, computational fluid dynamics (CFD) was

used and the rising bubble motion and its role for pyrolysis reaction were studied.
PT

They concluded that the heat transfer is mainly governed by the bubbles’ motions
CE

[248]. Xiong et al. studied the impact of bubbling bed hydrodynamics on temporal

variations in the exit tar yield for biomass fast pyrolysis using computational
AC

simulations of an experimental laboratory-scale reactor. They observed that the

trends in tar yield variability were complex and difficult to predict. In addition, the

standard deviation in tar yield enhanced by the increases in initial bed height in freely

bubbling state and became the maximum in slugging state [249]. Geometrical

optimisation of a fast pyrolysis bubbling fluidised bed reactor was studied by

51
ACCEPTED MANUSCRIPT

Papadikis et al. To optimise the condition, they used CFD. The reactor in Aston

University with the capacity of 1 kg/h was selected. The three-dimensional simulation

of the current operation of the reactor showed that a stationary bubble was formed

next to the feeding tube. Moreover, the size of the permanent bubble reached up to

the splash zone of the reactor [250].

T
IP
6.2. Circulating fluidised beds and transported beds

CR
Circulating fluidised bed (CFB) and transported bed reactor system’s operation is

similar to bubbling beds described above except that the residence time for the char

US
is almost the same as for vapours and gas. In addition, comparing to the fluidised

bed reactor, the char content in the collected bio-oil could be higher. The advantage
AN
of CFB is that it is suitable for very large throughputs even though the

hydrodynamics are more complex. Because of this features, it is widely used at very
M

high throughputs in the petroleum and petrochemical industry. From the recirculation
ED

of heated sand from a secondary char combustor, the heat is supplied [3,4]. A

common schematic of circulating fluidised bed reactor is shown in Fig. 23.


PT

A plan with the capacity of 650 kg/h was built by Ensyn [251,252]. In Canada and
CE

USA, circulating fluidised bed reactor was used to produce flavours for food with the

capacity of 2000 and 1700 kg/h [251,253]. Recently, companies like Metso, UPM
AC

and Fortum developed a circulating fluidised bed reactor with the capacity of 400

kg/h [251].

Velden et al. studied the pyrolysis of biomass in a circulating fluidised bed reactor.

They observed that CFB biomass pyrolysis produces mostly bio-oil. It’s yield at

500 °C exceeded 60 wt% [254]. Biomass pyrolysis in a circulating fluidised bed

52
ACCEPTED MANUSCRIPT

reactor for the production of fuels and chemicals was performed by Lappas et al.

They studied also the influence of adding ZSM-5 catalyst. The results showed that

the yield for bio-oil produced was 70 wt% without catalyst. They also observed that

catalytic biomass pyrolysis produced more water, coke and gases compared to

conventional pyrolysis. However, the bio-oil quality and composition was improved

T
[255]. One dimensional steady-state circulating fluidised-bed reactor model was

IP
developed by Trendewicz et al. In their simulations the feeding rate considered 0.023

CR
kg/s and also four different biomass feedstocks (pine, wheat straw, olive husks,

waste) were selected. They proved that pyrolysis reactions proceed quickly to 99%

US
conversion after 0.9 s of residence time. In addition, the highest organics yield of

65.5 wt% was obtained from pine, followed by 61.2 wt% from waste, 59.6 wt% from
AN
straw and 58.0 wt% from olive husks. They also concluded that the reactor model

provided an acceptable approximation of fluid dynamics, heat transfer and pyrolysis


M

product yields [256]. Dai et al. studied the pyrolysis of wood in a CFB reactor. They
ED

varied bed temperature, particle size of wood powder and the feeder position to

investigate their effects. They found out that the higher temperature and longer
PT

residence time increased the rate of the secondary reactions, producing fewer liquids.
CE

Moreover, the lower heating rate favoured the carbonisation and also declined the

liquid production [257].


AC

6.3. Rotating cone

Rotating cone reactor was developed in University of Twente for flash pyrolysis of

biomass [4]. The design of the rotating cone reactor is based on the intense mixing

of biomass and hot inert particles, which is the most effective way to transfer heat to

53
ACCEPTED MANUSCRIPT

the biomass. In comparison, the heat transfer of fluidised bed requires too much

ineffective inert carrier gas. Rotating cone reactor was developed, which required no

inert gases while simplifying the reactor parts and peripheral equipment as oil

condenser, gas cleaning etc. The original design of rotating cone reactor in 1989

relyed on a merely ablative principle, and no inert sand was used. Later the sand

T
transported-bed rotating cone reactor (RCR) was developed. After pyrolysis, the

IP
sand and char are further transported to a separate fluidised bed where the

CR
combustion of char takes place [3,4]. A schematic of the rotating cone reactor is

depicted in Fig. 24.

US
Junsheng pyrolysed wood shatters in a rotating cone reactor. The pyrolysis

temperature 550 °C and also vacuum pressure 0.08 MPa were used. In the optimum
AN
condition, the yield of bio-oil was 54.83 wt%. Methanol, ethanol, 1,3-Propanediol,

normal hexane, toluene, phenol cyclopentane and naphthalene were measured in


M

the bio-oil produced [258]. Wagenaar et al. studied the pyrolysis of wood in a rotating
ED

cone reactor. The effect of process parameters like the cone rotational speed (6–15

Hz), the reactor volume (3–200 L), the wood-dust feed rate (1–3.5 g/s) and the
PT

reactor temperature (550−700 °C) on the product composition was investigated.


CE

They compared the experimental results with predictions of an integrated reactor

model. The differences between the results from the model and experiment in the
AC

measured weight fractions of gas, tar and char produced were always less than 10%.

Under the optimum reactor working conditions, the bio-oil yield was 70 wt% [259].

54
ACCEPTED MANUSCRIPT

6.4. Ablative pyrolysis

Ablative reactor is used for the production of bio-oil in high yield with reduced

equipment size, costs and improved controllability. It relies on the heat transferring

when biomass slides over a hot surface (Fig. 25). The reaction limitation in ablative

T
reactor is the rate of heat transfer through the biomass particle. Pressure, the

IP
relative velocity of wood on the heat exchange surface and the temperature are

CR
controlling the rate of reaction. The key features of the pyrolysis in an ablative

reactor are high pressure of particle on hot reactor wall, high relative motion between

US
particle and reactor wall and reactor wall temperature. In addition, no inert gas is

required [4].
AN
Mainly CNRS laboratories in Nancy, France worked on ablative pyrolysis reactors

[260]. The National Renewable Energy Laboratory produced bio-oil with the yield of
M

60–65 wt% using an ablative reactor [4,251]. Researchers in Aston University


ED

produced bio-oil with the yield of 70–75 wt% [261]. In Germany, an ablative reactor

with feeding rate 6 t/d was built and its bio-oil was used in an engine for power
PT

generation [4].
CE

Pyrolysis of whole wood chips and rods in a novel ablative reactor was

investigated by Luo et al. Their study showed that the bio-oil yield from fast pyrolysis
AC

of wood chips (10–20 mm) and wood crumbles (2–2 mm) were similar (60 wt%).

Moreover, the yield and composition of bio-oil for both ablative and fluidised bed

reactors were in the same range for biomass particles with the size <1 mm [262].

Gomez-Onedero et al. studied the characterisation of the bio-oil produced by the

pyrolysis of different biomasses in an ablative reactor. Their feedstocks were

55
ACCEPTED MANUSCRIPT

eucalyptus chips, camelina straw pellets and wheat straw pellets. The experiments

were carried out under atmosphere pressure at 550 °C. The bio-oils with yielding

42.4, 48.8, and 41.0 wt% for eucalyptus, camelina straw pellets and wheat straw

pellets were produced, respectively. The products’ compositions changed by both

feedstock type and the operating conditions [263].

T
IP
6.5. Grinding pyrolysis

CR
A schematic for grinding pyrolysis technology is shown in Fig. 26. Inside the

grinding pyrolysis reactor, the brittle outer layer of biomass is removed by the

US
abrasion force of the grinding activity, which is performed by the hot steal balls with

different sizes filled inside the reactor. The brittle layer will form char and the
AN
pyrolysis will continue with remained particle. The products from pyrolysis are

pushed outside the reactor by the inert gas flow and also by volatiles and non-
M

condensable gases produced during the pyrolysis. Char is also separated before the
ED

volatiles travel to the condensation system [45].

Grinding pyrolysis was developed recently in Australia with feeding rate of 10


PT

kg/h [264]. Hasan et al. proved that grinding pyrolysis has the capability of feeding
CE

particles with a wide range of particle sizes, which could be simultaneously pyrolysed

and grinded. In addition, there was no requirement for a fluidising gas and simplified
AC

the requirement for cooling and condensing the pyrolysis products [45].

6.5. Auger reactor

This type of reactor is made of an oxygen free cylindrical heated tube in which

augers move the biomass through. A schematic of auger reactor is shown in Fig. 27.

56
ACCEPTED MANUSCRIPT

The temperature inside the tube raises the feedstock temperature to the pyrolysis

one. A condenser is considered to liquefy the volatiles. The bio-oil and char yields for

the products of auger reactor are similar to the fluidised bed reactor yields [251,265].

Auger reactors are divided into four categories: single screw laboratory-scale reactor

(<1 kg/h), single screw laboratory-scale reactors with a large capacity or pilot-scale

T
(1–15 kg/h), single screw industrial-scale reactors (>15 kg/h) and twin-screw reactors.

IP
ABRI-Tech in Canada is the main company working on auger reactors. Some

CR
universities such as Auburn University (USA), KIT (FZK) (Germany), Mississippi

State University (USA), Michigan State University (USA), Texas A&M (USA) and

US
Washington State University (USA) are also studying the pyrolysis of biomass with

auger reactor.
AN
Yu et al. studied the pyrolysis of rice husks and corn stalk at temperatures of 350,

400, 450, 500, 550, and 600 °C. The solid residence time was 60 s. The optimum
M

temperature was 500 °C for both feedstocks (bio-oil yield: 51 wt% for rice husk and
ED

54 wt% for corn stalk) [266]. Roux et al. measured the bio-oil and char yield by

pyrolysing the wood in auger reactor. The yields of bio-oil and char obtained were
PT

56.1 wt% and 19.8 wt%, respectively [267]. Potato peel waste pyrolysis in an auger
CE

reactor was investigated by Liang et al. They produced bio-oil and char with the

yields of 22.7 and 30.5 wt% [268]. Garcia-Perez et al. developed a lab scale auger
AC

reactor to pyrolyse pine pellets. Bio-oil and char yields obtained were 57.8% and

30%, respectively [269]. Liaw et al. used an auger reactor to produce bio-oil (59 wt%

yield) and char (19 wt% yield) by pyrolysing wood at 500 °C [270]. Ingram et al.

pyrolysed oak and pine wood in a lab scale (1 kg/h) auger reactor at 450 °C. The

highest yields for char and bio-oil were 27.8 and 56.3 wt%, respectively [271]. The

57
ACCEPTED MANUSCRIPT

auger reactor was used for wood chips pyrolysis by Puy et al. The highest yield of

bio-oil (57 wt%) was obtained at temperature of 500 °C with the residence times of

180 and 300 s. On the other hand, char yield was the maximum (41.5 wt%) at 500 °C

and 90 s residence time [272]. Kim et al. studied the char yield from switchgrass and

pine wood pyrolysis at different temperature in an auger reactor (5 kg/h feeding rate).

T
They could produce the maximum char yield at 450 °C (31.3 and 26.6 wt% for

IP
switchgrass and pine wood) [273]. Pyrolysis of beech wood chips (2.0–4.5 mm) in

CR
an auger reactor was performed at 450 °C for a residence time of 10 min by

Morgano et al. The maximum bio-oil yield (48.8 wt%) with char yield 20.6 wt% was

US
achieved in their study [274]. Ferreria et al. used a negative inside the auger reactor.
AN
Their feedstock was medium density fibreboard. They could produce the maximum

bio-oil yield (40 wt%) at 600°C with the char yield 17.3 wt% at the same conditions.
M

The highest char yield (39.7 wt%) was obtained at 450 °C [275]. An industrial scale

auger reactor was tested by Azargohar et al. Wheat straw, sawdust, flax straw and
ED

poultry manure (particle size from 0.42 to 3.36 mm) were the feedstocks. The
PT

process temperatures of 400, 475 and 550 °C with a solid residence time of 15 min

were chosen. The maximum bio-oil yield was obtained from the sawdust at 475 °C
CE

(52 wt%) and the minimum bio-oil produced from wheat straw at 400 °C (9.2 wt%)

[276]. Brown et al. developed an auger reactor using two screws to prevent the
AC

reactor from clogging. The maximum bio-oil yield (>70 wt%) with minimum char yield

was obtained at a high flow rate of sweep gas (3.5 L min-1), high heat carrier

temperature (600 °C), high auger speeds (63 rpm) [277]. Raffelt et al. tested the

sawdust and wheat straw pyrolysis at 500 °C with the solid residence time 10 to 60 s.

58
ACCEPTED MANUSCRIPT

The bio-oil and char yields of 70 wt% and 14–18 wt% were obtained. Under the

same conditions, the bio-oil and char with the yields of 25–30 and 50–55 wt% were

achieved from the pyrolysis of wheat straw [278]. The production of engineered char

in a vertical auger pyrolysis reactor for carbon sequestration was investigated by

Brassard et al. They concluded that the desired properties of char could be achieved

T
by the optimisation of reaction conditions [279]. The optimisation of the process

IP
conditions in an auger reactor was studied by Mathew et al. For this aim a 500 L

CR
reactor and biomasses including Mesquite and rice straw were used. The maximum

bio-oil yield (42.6 wt%) was obtained at 500 °C and 30 rpm for mesquite sawdust

US
while for rice straw the bio-oil yield was 34.6 wt% at the optimum conditions (475 °C

and 50 rpm) [280]. Thangalazhy-Gopakumar et al. studied the physiochemical


AN
properties of bio-oil produced at various temperatures from pine wood using an

auger reactor. They proved that the quality of the bio-oil and the yield are highly
M

dependent on process parameters such as temperature, feedstock, moisture content


ED

and residence time [281].


PT

6.6. Fixed-bed reactor


CE

Fixed bed reactors are applicable more in lab scale experiments due to their

simplicity and easier controlling system. However, in large scales, its application
AC

needs further studied, especially the homogeneous temperature distribution inside

the reactor is required further investigation. Fixed-bed reactor was used by Xiong et

al. for the pyrolysis of rice husk biomass. In their study, temperature varied from 300

up to 800 °C to trace the coke formation. The heating rate was also changed. Their

results indicated that the coke formed at slow heating rate and lower temperatures

59
ACCEPTED MANUSCRIPT

was imporous with a smooth surface. However, by increasing the heating rate or

temperature it became porous. This was because of the more free radicals

formtation at the higher temperature or higher heating rate, which promoted the

condensation reaction of large molecules, especially aromatics [347].

In Table 12, the advantages and disadvantages of different types of reactor used

T
in biomass pyrolysis is summarised [4,251].

IP
CR
7. Catalytic pyrolysis

The bio-oil from fast pyrolysis cannot be used directly as an engine fuel and

US
needs upgrading. As an example, the chemical composition of bio-oil from wood

pyrolysis is shown in Fig. 28. Several upgrading methods such as catalytic


AN
hydrodeoxygeneation are used to improve the fuel quality of bio-oil (lower oxygen

content). However, fast catalyst deactivation and also the costs of the process
M

prevent their use, especially in commercial scale application [5–7]. To tackle this
ED

problem, catalyst is directly used in pyrolysis step to remove the oxygen in some

oxygenates. The catalyst can definitely influence the yields of pyrolysis products. In
PT

addition, the yields of products from the pyrolysis (non-catalytic and also catalytic) of
CE

rapeseed cake using different catalysts are presented in Table 13 [282]. In Table 14,

the comparison of the bio-oil produced yield from different feedstocks with and
AC

without using the catalyst is shown [283,284]. From Table 14 and also the results of

the other research groups, is can be concluded that the main product of the pyrolysis

in the liquid and gas phases are aromatics, CO2 and H2O [285-289].

60
ACCEPTED MANUSCRIPT

To clarify the effect of the catalysts on the products changes, the reaction

pathway of catalytic pyrolysis was studied. For this aim, ZSM-5 catalyst in three

types of the reactors including bench scale bubbling fluidised bed, fixed bed and

semi-batch pyroprobe reactors were chosen and the pyrolysis of wood sawdust and

some model compounds was performed. The process was continuous at low

T
biomass weight hourly space velocities (less than 0.5 h -1) and high temperature

IP
(600 °C). Their results showed that cellulose is converted into anhydrosugars (mainly

CR
levoglucosan). In addition, light compounds such as acetol and glycolaldehyde, gas,

char and coke were also formed. The coke formation is related to the

US
polycondensation of oxygenates and polycyclic aromatics which are produced during

degradation of cellulose [290,291]. The general reaction pathway of catalytic


AN
pyrolysis of cellulose is shown in Fig. 29 [291]. More detailed model compound

studies also were performed by other research groups as below.


M

The catalytic pyrolysis of lignin in pyrolysis-GC-MS systems (Pt-filament pulse


ED

pyrolyser and coil-type reactors) was studied by several research groups. For this

aim, the temperature in the range of 500-700°C was chosen. HZSM-5, Y-Zeolite and
PT

Pd/C were selected as a catalyst. The results showed that the pyrolysis of lignin
CE

produces low liquid and gas yields and high char yields. In the catalytic pyrolysis,

initially, lignin was decomposed into oxygenates such as phenols through the
AC

cleavage of β-O-4, α-O-4 linkages and other C O/C C bonds. In the next step, the

polycondensation of lignin and the repolymerisation of phenols occurred to form char.

The oxygenates produced could diffuse into the pore of the zeolite catalyst. As a

result aromatics and olefins were produced through a series of dehydration,

decarboxylation, decarbonylation, and oligomerization reactions. The usage of

61
ACCEPTED MANUSCRIPT

catalyst in lignin pyrolysis increased the water yield and decreased the organic liquid

yield. This is because of the increasing the dehydration, cracking and other reactions’

rates [291–294]. The catalyst effect on lignin pyrolysis has been studied by some

researchers. Their results showed that toluene, benzene and xylene were the major

liquid components. In addition, the highest liquid yield was 43% and the coke plus

T
char yields varied in the range of 15%–50% [291–294].

IP
Hemicellulose is a group of polysaccharides found in plant cell walls and can be

CR
extracted from different plant sources with heterogeneity. Many researchers chose

xylan as a model compound of hemicellulose in catalytic pyrolysis due to its high

US
content in hemicellulose. At 220–315 °C, the decomposition of xylan occurs and

converts into mainly CO2. The main products from the catalytic pyrolysis are bio-oil,
AN
gas and char [291,292]. In the bio-oil, furans are the most abundant compounds

[295,296]. The general reaction pathway of catalytic pyrolysis of hemicellulose and


M

lignin is presented in Fig. 30 [291]. Considering the main three components


ED

(cellulose, hemicellulose and lignin) together, biomass catalytic pyrolysis chemistry is

shown in Fig. 31 [297]. In addition, the aromatics selectivity of several bio-oils


PT

produced from different model compounds and biomasses catalytic pyrolysis (ZSM-5
CE

catalyst) are also presented in Table 15 [291].

Catalyst for pyrolysis of biomass can be packed/co-fed with the feedstock in the
AC

reactor (in situ configuration) or the catalyst can be set in the outlet of reactor to

upgrade the pyrolysis volatile gas (ex situ configuration) [298]. Both in situ and ex

situ convert the oxygenated compounds into stable products.

During in situ process, the catalyst helps the thermal cracking inside the reactor.

This can increase the light component content in bio-oil and reduce char content.

62
ACCEPTED MANUSCRIPT

The residence time in situ pyrolysis is 1–2 s, which is short. Therefore, to increase

the rate of cracking, higher amount of catalyst is needed. In addition, the coke

formation is high [298].

Comparing with the in situ catalytic pyrolysis, ex situ pyrolysis is more flexible and

the selectivity of the compounds in the products can be easily controlled. This is

T
because the catalyst is located in the outlet of reactor and can be easily separated

IP
[298].

CR
In situ and ex situ catalytic pyrolysis of pine wood with ZSM-5 catalyst was

compared by Yildiz et al. Their results showed a decrease (2.3 wt%) of char yield for

US
in situ process while 0.2 wt% decrease in ex situ process was observed. In addition,

gas yield increase was also different (4 wt% for in situ and 1.7 wt% for ex situ). The
AN
bio-oil yield of 17 wt% was obtained for both processes. However, the aromatics

content produced in the in situ process was higher [299]. Other researchers studied
M

the effect of in situ and ex situ pyrolysis on the removal of oxygen. Their results
ED

indicated that ex situ process remove more oxygen from biomass and the bio-oil

produced had lower oxygen content. This can be due to the less homogeneity of
PT

catalyst [300,301].
CE

The catalysts used for pyrolysis are mainly solid acids such as zeolites, silica–

alumina, silicalite, FCC catalysts, alumina, molecular sieves, as well as metal oxides
AC

such as zinc oxide, zirconia, ceria, and copper chromite. In addition, other inorganic

materials including metal chlorides, phosphates, sulphates, and alkali metals have

also been studied in catalytic pyrolysis [286,302–304]. The catalysts can be divided

into two main groups, as detailed in the following section.

63
ACCEPTED MANUSCRIPT

The catalyst type and its structure can effectively influence the products yields

and compositions (Fig. 32). For instance, the yield of bio-oil was reported higher for

the ZSM-5 comparing HY and USY catalysts using the same biomass feed and

process conditions [297]. Furthermore, ZSM-5 in the same research produced higher

contents of the aromatics. Therefore, the catalyst development is a key parameter to

T
control the pyrolysis process and consequently the products yields and compositions.

IP
The suitable catalyst should catalyse the desired reactions such as dehydration,

CR
hydrogenation, decarbonylation, decarboxylation, C–C coupling and cracking. So far,

from the used catalysts for the biomass pyrolysis, zeolite based catalysts because of

US
having longer activity time and also producing more stable bio-oil are the most

popular catalyst.
AN

7.1. Zeolite catalysts for CFP of biomass


M

Zeolite catalysts are the commonly used catalysts in the catalytic pyrolysis of
ED

biomass. These include ZSM-5, Y zeolite, Beta zeolite, MCM-41, CM-22, Mordenite,

SAPO-34 [305]. In Table 16, a summary of different zeolite catalysts used for
PT

catalytic pyrolysis under different conditions is shown [64]. Several research groups
CE

worked on different zeolite catalysts. Their results indicated that the zeolite catalyst

increases the aromatic yields, and consequently promoted coke formation [287]. The
AC

selectivity between desired aromatics and coke is related to the pore size of the

catalyst. The bigger pore size favours the formation more coke [288]. The other

important parameter of the catalyst is the acidity. For instance, comparing with

zeolite, silicalite catalyst without acidity but have the same pore architecture

produces less aromatics and more coke [291]. Modification of zeolite catalyst with

64
ACCEPTED MANUSCRIPT

metals also has been studied. For this aim, metals such as Ga, Mo, Co, Ni, Fe, Zn,

Pt or Pd were used. The result indicated that metal can effectively promote the

aromatic yield, especially Ga [291,306–308]. In another research, mesoporous

zeolite including MCM-41, Cu-Al-MCM-41SBA-15 Al-MCM-41, Al-SBA-15 were used

for catalytic pyrolysis of biomass. Their results showed that mesoporous zeolite

T
influences the composition of bio-oil in a way that phenolics’ yields were enhanced

IP
while the carboxylic acid and carbonyl yields were decreased [309,310]. The

CR
deactivation of zeolite catalyst is also studied by different groups. Deactivation was

mainly caused by coke formation. Normally coke deactivates the catalyst by

US
depositing on the surface (thermal coke) and filling the pores (catalytic coke). From

crosslinking and polycondensation of biomass and secondary reaction of pyrolysis


AN
vapors including aldol condensation and Diels–Alder reactions, the thermal coke

forms while catalytic coke is caused by hydrogen transfer and carbonium ion
M

chemistry [311].
ED

ZSM-5 is one of the most used catalysts in the literature to catalyse the pyrolysis

of biomass. In Fig. 33, the chemical pathway of hollocellulose conversion over ZSM-
PT

5 catalyst is presented [305]. Comparing with ZSM-5, HZSM-5 has stronger acidity,
CE

shape-selective with a good thermal and hydrothermal stability that favours catalytic

pyrolysis of biomass than ZSM-5. The acidic sites of HZSM-5 promote


AC

deoxygenation, decarbonylation and decarboxylation reactions of oxygenate

components. In addition, it improves the cracking, oligomerisation, alkylation,

isomerisation, cyclisation and aromatisation reactions via carbonium ion

mechanisation. HZSM-5 is the most effective catalyst for the production of aromatics

from the biomass pyrolysis vapour [286,312]. Pattiya et al. studied the catalytic

65
ACCEPTED MANUSCRIPT

pyrolysis of cassava rhizome. They observed that the catalyst improved the

production of aromatics while it reduced the formation of oxygenated lignin

derivatives [313]. Williams et al. investigated the pyrolysis of rice husks over ZSM-5

catalyst. They observed that using catalyst significantly decreased the yield of bio-oil

and also oxygen content of the oil. Severe coke formation on the catalyst was also

T
observed [314]. Microalgae was pyrolysed using HZSM-5 catalyst in a fixed bed

IP
reactor by Pan et al. They concluded that catalyst resulted in bio-oil with low oxygen

CR
content (19.5 wt%) and high heating value (32.7 MJ/kg) comparing with that from

non-catalytic pyrolysis [315]. Zhang et al. used HZSM-5 catalyst to pyrolyse

US
corncobs in a fluidised bed rector. They also confirmed the decline of oxygen content

and the increase of heating value for the bio-oil produced (25% oxygen content
AN
decrease and high heating value of 34.6 MJ/kg) [316]. The catalytic pyrolysis of

beech wood using HZSM-5 was carried out by Stephanidis et al. Their results
M

showed an increased content of aromatics in the bio-oil produced comparing with


ED

that of the non-catalytic pyrolysis [141]. Foster et al. studied the catalytic pyrolysis of

glucose, furan, and maple wood over ZSM-5 catalyst in a fixed bed reactor. They
PT

observed that by increasing the ratio of silica to alumina in the catalyst, the aromatics
CE

contents altered and at the ratio of 30, the aromatics content in the bio-oil produced

reached the maximum [8]. The effect of different zeolite catalysts (HZSM-5, Fe-ZSM-
AC

5 and H-Beta) on the bio-oil yield produced from the pyrolysis of rapeseed cake was

carried out by Giannakopoulou et al. At 400 °C, they could achieve the bio-oil yields

of 63.0, 59.4 and 58.8 wt% for H-Beta, HZSM-5 and Fe-ZSM-5, respectively [318].

Beta, Y, ZSM-S, and mordenite were used as a catalyst by Aho et al. in the pyrolysis

of pine. The yield and especially composition of the bio-oils produced were different.

66
ACCEPTED MANUSCRIPT

ZSM-S resulted in the formation of more ketones [319]. Catalytic pyrolysis of

biomass with AI-MCM-41 and siliceous MCM-41 catalysts was carried out by

Iliopoulou et al. and the results were compared with non-catalytic pyrolysis. The

yields and compositions of products were different. AI-MCM-41 catalytst enhanced

the phenolics concentration while it reduced the formation of the acids. AI-MCM-41

T
with high Si/Al ratio enhanced the organic components yield in the bio-oil. However,

IP
lower Si/Al ratios increased the yields of gas and coke [309]. Adam et al. tested AI-

CR
MCM-41 catalysts for the pyrolysis of spruce wood. Four different AI-MCM-41 type of

catalysts with a Si/ Al ratio of 20 were used. Cu cations were also added into the

US
structure. Their results showed that the use of catalyst resulted in elimination of

levoglucosan. In addition, the catalysts increased the yield of acetic acid and furan
AN
[320]. Li et al. chose ZSM-5 catalyst for biomass pyrolysis. They observed that the

catalyst resulted in the formation of more aromatics. They also modified ZSM-5 with
M

the metals such as Ga, Mo, Co, Ni, Fe, Zn, Pt and Pd and observed that some metal
ED

addition promoted the aromatic yield, especially Ga [321]. Chen et al. proved that

modifying ZSM-5 catalyst enhanced the aromatics yield [322]. Engtrakul et al. related
PT

the acidity of ZSM-5 with the formation of aromatics during the pyrolysis of pine
CE

wood. They concluded that with the increase of the acidity of zeolite, the content of

alkylated aromatics decreased. This could be related to the change of the cyclization
AC

and alkylation reactions [317].

7.2. Non-zeolite catalysts for CFP of biomass

Several non-zeolitic catalysts were also tested in catalytic pyrolysis of biomass.

The main groups include metal oxides and noble or transition metals. In a research,

67
ACCEPTED MANUSCRIPT

Na2CO3/Al2O3 catalyst was used in biomass pyrolysis. The result indicated that the

bio-oil produced was very instable. Platinum on the surface of the Na2CO3/Al2O3 was

planted to improve the quality of bio-oil, which, however, did not impact much the

stability of bio-oil. Two catalysts containing Na2CO3/Al2O3 and Pt/Al2O3

simultaneously were used to improve the bio-oil quality. This modification could

T
significantly improve the stability of bio-oil [323]. Cu, Fe, and Zn on aluminium oxide

IP
support were also tested in the catalytic pyrolysis of biomass. These catalysts

CR
favoured the conversion of oxygenates to aromatic and aliphatic hydrocarbons

[324,325]. Other catalysts such as TiO2, CeO2, CeOx-TiO2, ZrO2 and MgO were also

US
tested. This group of catalysts had a high tendency for promoting ketones production

[326]. Ni/Al2O3, Ni/ZrO2, Ni/CeO2 and Ni/SiO2 were used in catalytic pyrolysis of
AN
biomass. These catalysts could enhance the production of bio-oil with a high yield. It

was observed that Ni/CeO2 and Ni/SiO2 produced less valuable compounds [327]. In
M

Table 17, a summary of non-zeolite condition experiments is presented [328].


ED

Nguyen et al. [28] studied the use of a Na2CO3/Al2O3 catalyst in biomass catalytic

pyrolysis. They proved the decrease of oxygen content in bio-oil and also its heating
PT

value. However, the bio-oil produced was very unstable. Therefore, a dual-bed
CE

system consisting of two separate units using Na2CO3/Al2O3 and Pt/Al2O3 catalysts

was proposed for improving the stability of bio-oil. The catalysts used could help to
AC

significantly improve the bio-oil stability [323]. Mineral catalysts such as dolomite,

limestone, ZnO and MgO were used by Nokkosmaki et al. to pyrolysis pine sawdust.

Comparing non-catalytic pyrolysis, the carbon yields in catalytic pyrolysis decreased.

ZnO catalyst could not influence the products composition significantly comparing to

non-catalytic pyrolysis. However, MgO, dolomite and limestone produced more

68
ACCEPTED MANUSCRIPT

hydrocarbons in bio-oil [329]. Dilcio et al. carried out catalytic pyrolysis by using

colloidal FeS catalyst. The catalyst could influence the yields and compositions of

the products to some extent. Moreover, the addition of a dispersed iron sulphide

catalyst could decline the oxygen content in the bio-oil by a further 10 wt% [330].

Transition metals were also tested for biomass catalytic pyrolysis. Dilcio et al. tested

T
the commercial sulphided Ni/Mo/Al2O3 catalyst at 400 °C. They could decrease the

IP
oxygen content of the bio-oil from 20 to 10 wt% [330]. Dickerson et al. summarised

CR
that base catalysts influence in the properties of bio-oil produced. Their review

showed that the bio-oil yield and its oxygen content decreased with the use of base

US
catalysts [285]. Catalytic pyrolysis of cotton seed was performed at a different

loading of MgO catalyst. The catalyst influenced the pyrolysis reaction and
AN
decreased the oxygen content in bio-oil from 9.56% to 4.90% [331]. Barbooti et al.

pyrolysed reed with 0.6–10 wt% potassium carbonate (K2CO3). Their results showed
M

that K2CO3 significantly changed the composition of the gas during the pyrolysis.
ED

Comparing non-catalytic pyrolysis, higher content of carbon monoxide and hydrogen

was produced [332]. Red mud catalysts like red brick (calcined red mud) were used
PT

by Gungor et al. in a fixed bed reactor to pyrolyse pine wood. The results showed
CE

that the bio-oil contained substantial amount of phenolics and sugar [301]. Recently,

an iron based catalyst was used by Moud et al. to upgrade the pyrolysis gas
AC

produced from chips of treetops and branches at 450 °C. Their results proved that at

a space velocity of 1100 h−1 for 8 h, no significant catalyst deactivation was observed.

In addition, 70%–80% reduction of acetic acid, methoxy phenols and catechol and a

55%–65% reduction of non-aromatic ketones, benzene, toluene, xylene and

heterocycles were measured. All the other compounds yields were also influenced

69
ACCEPTED MANUSCRIPT

by using this catalyst. This could prove that iron based catalyst could be a potential

candidate for upgrading the pyrolysis gas [348]. Beech wood catalytic pyrolysis was

investigated by Stummann et al. to pre-upgrade the bio-oil prior to hydrotreatment

stage. Their experiment was performed in a fluidised bed reactor using

CoMoS/MgAl2O4 catalyst. They also studied the influence of temperature and

T
pressure on the products’ yields and compositions. Their results showed that higher

IP
temperature produced more gas and less char. The pressure increase enhanced the

CR
bio-oil yield while the CO and CO2 yields decreased. They could conclude that

catalytic pyrolysis could be a potential process to produce the fuels could replace the

US
fossil fuels [349]. Zhang et al. studied the effects of metal oxide catalyst including

Al2O3, SiO2, ZnO, K2O, MgO, CaO, La2O3 on the pyrolysis of poplar wood, cellulose
AN
and lignin. For this aim, they used a fixed bed reactor at 500°C. They could show

that the basic catalysts such as CaO and MgO promoted the formation of gaseous
M

products. However, the acidic oxides produced more tar. On the other hand, the
ED

basic oxides suppressed tar formation. In addition, the basic catalysts like CaO

could suppress the production of phenolic compounds. Basic oxides catalysts also
PT

resulted in more coke formation [350].


CE

Comparing with non-catalytic pyrolysis, the catalytic pyrolysis can improve the

quality of the bio-oil produced and also decrease the oxygen content in the bio-oil.
AC

Selecting the correct process conditions and an appropriate catalyst is the key to

enhance the properties of bio-oil. The reactions involved in the catalytic pyrolysis are

very complex. Therefore, the design of a catalyst with good stability and activity is

essential to commercialise the process. Zeolite catalysts are capable of catalysing

the effective upgrading of bio-oil during catalytic pyrolysis of biomass, and the cost

70
ACCEPTED MANUSCRIPT

for using them is relatively low. However, Zeolite-based catalysts suffer from early

coke formation, resulting in the rapid deactivation of the catalyst. The formation of

coke consumes the organics in bio-oil and significantly shortens the working life of

the catalyst, leading to the lower yields of biofuels and the higher operating cost.

Further studies to optimise the catalytic pyrolysis process and the development of

T
more effective and robust catalysts are required.

IP
CR
8. Processes for biomass catalytic pyrolysis improvement

US
8.1. Co-catalytic fast pyrolysis of biomass with other feedstocks

Catalytic fast pyrolysis of biomass, even in the presence of highly efficient


AN
catalyst, produces a low yield of aromatic hydrocarbon with a high yield of coke. On

way to improve this is the addition of hydrogen-rich co-reactant such as waste


M

plastics or tire waste. It was proved by several groups that co-processing of biomass
ED

with of hydrogen-rich co-reactant can significantly improve the yield of aromatics and

lower the coke formation [333]. The Co-catalytic fast pyrolysis of biomass with
PT

plastics is normally performed at 400–700 °C using zeolite base catalysts [334].


CE

Fixed-bed reactors are the commonly used type for this aim. The reaction

mechanism in this process is relatively complex and can be categorised into two
AC

main groups: reactions during thermal degradation and reactions between produced

compounds. In Table 18, a summary of researches on the co-pyrolysis of biomass

with plastics and tier waste is presented [335].

71
ACCEPTED MANUSCRIPT

Alcohol was chosen to co-process with biomass. The results from different

groups showed that co-processing of different alcohols by biomass can produce

more aromatics and olefins in the bio-oil [336,337].

Kuppens et al. indicated that co-pyrolysis was more beneficial than pyrolysis and

it had the potential for commercialisation [328]. Akhtar et al. and Velghe et al.

T
concluded that the optimum temperature for biomass co-pyrolysis to obtain the

IP
maximum bio-oil yield (>45 wt%) was 400–600 °C [339,340]. Martinez et al. carried

CR
out the co-pyrolysis of biomass and waste tires. They performed the experiments in

two different reactors (fixed-bed reactor and auger reactor). They proved that the

US
auger reactor produced more bio-oil than the fixed-bed reactor [341]. Brebu et al.

studied the co-pyrolysis of pine cone with low density polyethylene (LDPE),
AN
polypropylene (PP), and polystyrene (PS). They conclude that by mixing the pine

cone and polymers in the same weight ratio, the bio-oil yield significantly increased.
M

In addition, energy content of bio-oil produced from mixed feedstock was higher than
ED

the bio-oil produced from pine cone [342].

From literature, many researchers proved that co-pyrolysis can significantly


PT

improve the quantity and quality of the bio-oil. They tried different materials including
CE

plastics and tire wastes with biomass and explained the process as a cheap way of

producing a sustainable fuel for replacing fossil fuels. The process was generally
AC

simple as no catalysts or solvents are required in the process. Using different wastes

in co-pyrolysis of biomass can also reduce the need for landfills. However, it should

be noted that co-pyrolysis process needs more researches to clarify the optimum

conditions for the process, most importantly, to understand the interaction of the

different intermediates produced from the different feedstock and how do the

72
ACCEPTED MANUSCRIPT

interaction affect the properties of the fuel produced. From literature, it can be also

found out that each biomass will nictitate own special optimum process conditions. In

conclusion, the co-pyrolysis still requires more studies to establish an appropriate

method in selecting and the ratio of material to mix with biomass.

T
8.2. Recycling of tail-gas

IP
Inert environment used in catalytic pyrolysis of biomass. Different gases were

CR
tested as an inert carrier gas such as H2, CO, CO2, CH4 and N2. The researchers

found that different carrier gases influence the yield and composition of the pyrolysis

US
products [205]. The tail-gas from the pyrolysis was recycled into the reactor. The

result indicated that bio-oil composition influence significantly by using tail-gas as the
AN
carrier gas. The bio-oil produced was rich in aromatics and phenols while the acids

and levoglucosan content declined [343,344].


M

Mullent et al. used a fluidised bed reactor with a recycled gas modification for the
ED

pyrolysis of white oak, switchgrass, and pennycress presscake. They compared the

results with those produced under inert N2. The recycle gas, which was mainly
PT

composed of CO, CO2, H2 and light hydrocarbons could increase the deoxygenation
CE

rate of bio-oil. The bio-oils produced under the recycle gas atmosphere included high

amount of aromatic hydrocarbons and nonmethoxylated phenolics. However, it had


AC

lower concentrations of levoglucosan and acids than the bio-oils in the N2

atmosphere [343]. Dorado et al. studied the co-processing of agricultural plastic

waste and switchgrass via tail gas reactive recycling. They observed that under an

atmosphere of approximately 70% recycled tail gas, the bio-oil produced was mainly

73
ACCEPTED MANUSCRIPT

composed of aromatics. When the atmosphere had a 55% of recycled tail gas, the

bio-oil yield increased while the oxygen content of bio-oil enhanced [344].

9. Progress in commercialisation of pyrolysis

Pyrolysis can be performed in a batch or continuous processes. In consideration

T
of the high cost of the batch processes and the low production rate, the continuous

IP
pyrolysis was always the interest of the researchers. So far, there have been much

CR
efforts done by scientists to commercialise the pyrolysis process. However, from the

open literature, it can be concluded that the pyrolysis technology is still in the early

US
stages. Although the complicated reaction network during pyrolysis is studied a lot,

the other aspects of pyrolysis such as controlling system of the reactor require
AN
further investigations. Different groups of scientists used their own design of reactor

to carry out the biomass pyrolysis. However, they are still experiencing difficulties in
M

scaling up the pyrolysis process. In addition, a number of companies are working on


ED

the commercialisation of the pyrolysis and several commercial plants in a moderate

size were built up. Table 19 summarized the pyrolysis plants of the commercial or
PT

demonstration size is shown [4,26,229,345].


CE

Ensyn, an American company, established in 1984 conducted the

commercialisation of pyrolysis of non-edible biomass for fuel production. From 1985


AC

to 1989, with the joining of Red Arrows company to Ensyn, further commercialisation

of pyrolysis process took place using the fluidised bed reactor. In 1996, a pyrolysis

plant with the feeding rate of 1667 kg/hr was built up by Red Arrows-Ensyn

companies. In their plant, the temperature of pyrolysis was 520 °C with the 1–2 s of

residence time. Pyrovac in Canada developed a vacuum pyrolysis process for

74
ACCEPTED MANUSCRIPT

conversion of biomass to liquids in 2000 with the feeding capacity of 3500 kg/h and

the pyrolysis temperature of 500 °C. Bio-oil was produced with the yield of 30–45

wt%, which was lower than the bio-oil yield reported with the fluidised bed reactors

(~70 wt%). This process required a high investment and maintenance costs with a

low heating efficiency. Consequently, further improvement is needed in the larger

T
scales. Genting and BTG companies completed the first pyrolysis unit in Malaysia in

IP
2005. Rotating cone reactor was used for this aim. The bio-oil yield of 60 wt% was

CR
achieved in their plant, which was lower than the bio-oil yield obtained from the

fluidised bed reactor. In 2013, Fortum/Valmet companies commercialised the

US
pyrolysis technology for the production the bio-oil with the feeding rate of biomass of

10000 kg/h in Finland. Fluidised bed reactor at 500 °C was used in their process. In
AN
2014, BTG-BTL/EMPYRO in Netherlands developed a rotating cone reactor to

pyrolyse the biomass with the capacity of 5000 kg/h. In their system, the heating was
M

rapid with a low resistance time of vapour in the reactor. This could decrease the
ED

costs of the process by some extent. However, still the bio-oil yield was lower than

the one from fluidised bed reactor. In the same year, KiOR company in USA could
PT

convert the biomass with the feeding rate of 21000 kg/h to produce bio-oil using a
CE

catalytic reactor system. Clays, metal oxides and alkaline earth metals were chosen

as the catalysts in their pyrolysis system. Nevertheless, the fuel properties of the bio-
AC

oil produced were improved significantly using this process, the catalyst deactivation

was recognised the major problem of the process. UPM company in Finland

produced biodiesel in a commercial size of biorefinery unit (11574 kg/h) in 2015. The

catalytic pyrolysis followed by hydrotreatment was used in their process. In 2017, AE

Cote-Nord Bioenergy/Ensyn companies used fluidised bed reactor to commercialise

75
ACCEPTED MANUSCRIPT

the biomass pyrolysis process with the feeding capacity of 9000 kg/h. The similar

results of other fluidised bed reactor trials were observed in their process

[4,26,229,345]. In addition, due to higher yield of bio-oil and easier operation,

fluidised bed reactor is mainly considered in the larger scales. From all the discussed

above, it could be concluded that there are a number of barriers that must be

T
overcome to advance the technology of biomass pyrolysis to commercialisation.

IP
CR
10. Economics and feasibility study of the pyrolysis

The biomass pyrolysis has a potential to be commercialised. The current

US
commercial plants currently are mainly in the moderate scale and further funding and

demands are required to commercialise them in a larger and widespread form. The
AN
economics and feasibility of the biomass pyrolysis plant was investigated by several

research groups. One of the initial complete studies was performed by Ringer et al.
M

in NREL in 2006 [345]. For this aim, they divided the pyrolysis process into several
ED

steps including feed handling and drying, pyrolysis, char combustion, product

recovery and steam generation. A model by Aspen software was developed for this
PT

aim. For a plant with 550 ton/day biomass (wood chips, 50% by mass water content)
CE

feed, the cost of the bio-oil from a pyrolysis plant including 10% internal rate of return

was calculated $7.62/GJ on a lower heating value (LHV) basis [345]. Recently, the
AC

economic aspect of the biomass pyrolysis plant was studied by

Jaroenkhasemmeesuk et al. For this aim, they used a demonstration size plant with

the feeding rate of 10–20 kg/h biomass. Their study showed that in a large scale, the

final price for the bio-oil is still high and not comparable with fossil fuel prices.

76
ACCEPTED MANUSCRIPT

However, they could also conclude that with further studies and process optimisation

it can be feasible [346].

All in all, still the price of bio-oil considering the upgrading expense is higher than

that fossil fuel prices and further process consideration is required to make it feasible.

T
11. Outlooks

IP
Pyrolysis is touted to be a thermochemical process to convert the biomass to an

CR
engine fuel. However, the costs, operation problems such as coke formation and

also the produced fuel quality is still limiting its use as a replacement of fossil fuels.

US
In this paper, we summarised the production of bio-oil through pyrolysis process

from the slow or fast pyrolysis to the catalytic pyrolysis. The effects of essential
AN
experimental parameters, the process design, the reactors and the catalysts on the

pyrolysis process were also summarized and discussed as well. Most of the
M

researches in this field are fundamental. The existing literature on pyrolysis


ED

modelling indicates that great efforts have been made in understanding of the

reaction pathway of the pyrolysis. However, Understanding of the kinetic of pyrolysis


PT

process is a central factor. To the best of our knowledge, a suitable kinetic and
CE

reaction mechanism pathway has not been yet proposed. Different types of biomass

pretreatments such as physical, thermal, chemical and biological methods were


AC

developed to enhance the bio-oil yield and quality. Most of the pretreatment

processes could increase the efficiency of pyrolysis. However, how do the

pretreatment impact the structural configuration of biomass feedstock and how do

the pretreament relate with the composition and properties of the bio-oil produced

needs further investigation. In addition, the effects of different process conditions like

77
ACCEPTED MANUSCRIPT

heating rate, temperature, biomass particle size were investigated. They all showed

a significant influence on yields and compositions of the products. The different

parameters impacted the pathways for the pyrolysis, the fundamental mechanism of

which needs further investigation, in order to obtain guidance or develop the

methodology for optimisation of the pyrolysis process both in lab scale or industrial

T
scale. Different reactor configurations were developed by different researchers to

IP
maximize the yields of bio-oil or to improve the properties or some specific properties

CR
of bio-oil. However, a large number of the reactor still cannot fully satisfy the

requirements the commercial reactors for the pyrolysis of biomass in industrial scale.

US
A pyrolyser to meet the requirement for commercial use should have the features of

processing versatile feedstock, low requirement of the feedstock (i.e. particles,


AN
moisture content), low requirement to reactor materials, high heat transfer efficiency,

high processing capacity and etc. to reduce the operating cost. The existing
M

knowledge in literature showed that catalytic pyrolysis of biomass either in situ or ex


ED

situ was developed to improve the property of bio-oil. However, many challenges

such as catalyst deactivation can still be addressed. The development of a suitable


PT

catalyst which can stand for an acceptable period of time in the commercial scale
CE

with a reasonable price is the key for the catalytic pyrolysis. The design of a catalyst

will definitely need of the better understanding of the biomass decomposition


AC

chemistry, which still requires more fundamental studies. So far different catalysts

were developed or tested for the catalytic fast pyrolysis of different biomasses.

Zeolite basis catalysts have emerged as a potentially promising one in converting

biomass to a fuel with a low oxygen content. However, they all still suffer from coke

formation and the resulting issue of rapid deactivation. Furthermore, the low yield of

78
ACCEPTED MANUSCRIPT

bio-oil and also carbon deficiency are the other parameters, which are required to be

considered in working with the zeolite basis catalysts. Co-pyrolysis of biomass with

other wastes such as plastics and tire waste was investigated by a number of

researchers. The different studies showed that the effective pyrolysis of different

biomasses required different conditions with an appropriate waste to be co-

T
processed. Therefore, the co-processing of biomass with different materials will

IP
require further fundamental studies about the different kinetics for pyrolysis of the

CR
different biomass, more importantly, the interaction of the intermediates produced

from the different biomass during the co-pyrolysis. This is because the interaction of

US
the reaction intermediates could determine the process efficiency of pyrolysis and

the composition and properties of the bio-oil produced. Concluding all, the
AN
economical production of a fuel from biomass with a comparable composition with

fossil fuels is still in the initial stage of the development. Further fundamental studies
M

about the pyrolysis process are required to fully understand the chemistry of the
ED

pyrolysis to provide a knowledge base for commercialising the pyrolysis process. To

commercialise the pyrolysis process, the following recommendations could be used.


PT

First of all, universally accepted specifications or standards for the bio-oil is required
CE

to be prepared. Secondly, the influence on the environment, safety and health of

production personnel potentially exposed to bio-oil is required to be established.


AC

Furthermore, some of bio-oil specifications such as odour make its use limited as it is

not accepted in the public. Such a problem should be solved. Currently, limited funds

are available for expanding the bio-oil production plants. In addition, the numbers of

bio-oil suppliers are limited. A secure source of biomass feed is required to

guarantee the production. In long term, the biomass limitation can influence the bio-

79
ACCEPTED MANUSCRIPT

oil production. The char separation from the bio-oil is another aspect, which should

be solved in the future studies. Although it is recognised by the biomass pyrolysis

community, there is not a consensus in how this can be done easily at a reasonable

cost. Because of low demand, a large scale biomass pyrolysis plant was not

established. All the commercial ones are currently in the moderate size. Therefore,

T
the appropriate advertisement is crucial to show its advantages. The techno-

IP
economical aspect of the pyrolysis shows that the upgrading of bio-oil is still

CR
requiring further studies.

Acknowledgments
US
The authors would like to thank University of Tabriz for their support. This work
AN
was also supported by the Strategic International Scientific and Technological

Innovation Cooperation Special Funds of National Key R&D Program of China (No.
M

2016YFE0204000), the Program for Taishan Scholars of Shandong Province


ED

Government, the Recruitment Program of Global Young Experts (Thousand Youth

Talents Plan), the Natural Science Foundation of Shandong Province


PT

(ZR2017BB002) and the Key R&D Program of Shandong Province


CE

(2018GSF116014).
AC

80
ACCEPTED MANUSCRIPT

References

[1] L.F. Cabeza, A. Palacios, S. Serrano, D. Urge-Vorsatz, C. Barreneche, ‎J. Renew.


Sustain. Energy, 27 (2018) 681-688..

[2] S.Wang, G. Dai, H. Yang, Z. Luo, Prog. Energy. Combust. Sci.,. 62 (2017) 33-86.

T
[3] D. Mohan, C.U. Pittman, P.H. Steele, Energy. Fuels., 20 (2006) 848-889.

IP
[4] A.V. Bridgwater, Biomass Bioenergy, 38 (2012) 68-94.

CR
[5] M. Gholizadeh, R. Gunawan, X. Hu, S. Kadarwati, R. Westerhof, W. Chaiwat,

US
M.M. Hasan, C.-Z. Li, Fuel Process. Technol., 150 (2016) 132-140.

[6] M. Gholizadeh, R. Gunawan, X. Hu, M.M. Hassan, S. Kersten, R. Westerhof, W.


AN
Chaitwat, C-Z. Li, Fuel Process. Technol., 146 (2016) 76-84.
M

[7] M. Gholizadeh, R. Gunawan, X. Hu, F.M. Mercader, R. Westerhof, W. Chaitwat,


M.M. Hasan, D. Mourant, C-Z. Li, Fuel Process. Technol.,148 (2016) 175-183.
ED

[8] A. J. Foster, J. Jae, Y.-T. Cheng, G.W. Huber, R. F. Lobo, Appl. Catal., A, , 154
(2012) 423– 424.
PT

[9] L. Sun, X. Zhang, L. Chen, B. Zhao, S. Yang, X. Xie, J. Anal. Appl. Pyrolysis, 121
CE

(2016) 342–346.
AC

[10] S. Dua, D.P. Gamlielb, J.A. Vallab, G.M. Bollasb, J. Anal. Appl. Pyrolysis, 122
(2016) 7–12.

[11] Y. Zheng, F. Wang, X. Yang, Y. Huang, C. Liu, Z. Zheng, J. Gua, J. Anal. Appl.
Pyrolysis, 126 (2017) 169–179.

81
ACCEPTED MANUSCRIPT

[12] R. French, S. Czernik, Fuel Process. Technol., 91 (2010) 25–32.

[13] F.X. Collard, J. Blin, Renew. Sust. Energy. Rev., 38 (2014) 594–608.

[14] D. Vamvuka, Bio-oil, Int. J. Energ. Res., 35 (2011) 835–62.

[15] J.P. Lange, , Bio. F. Pr.,1 (2007) 39–48.

T
IP
[16] Z. Chen, E. Leng, Y. Zhang, A. Zheng, Y. Peng, X. Gong, Y. Huang, Y. Qiao, J.

CR
Anal. Appl. Pyrolysis, 130 (2018) 350-357.

[17] M.V.D. Velden, J. Baeyens, A. Brems, B. Janssens, R. Dewil, Renew. Energ.,


35 (2010) 232–242.
US
AN
[18] M. Garcia-Perez, A. Chaala, H. Pakdel, D. Kretschmer, C. Roy, Biomass
Bioenergy, 31 (2007) 222-242.
M

[19] C.A. Mullen, A.A. Boateng, J. Anal. Appl. Pyrolysis, 90 (2011) 197-203.
ED

[20] M.C.B Lopez, C.G. Blanco, A. Martinez-Alonso, J.M.D. Tascon, J. Anal. Appl.
Pyrolysis, 65 (2011) 313-322.
PT

[21] P. Morf, P. Hasler, T. Nussbaumer, Fuel, 81 (2002) 843-853.


CE

[22] L. Zhang, X. Hu, K. Hu, C. Hu, Z. Zhang, Q. Liu, S. Hu, J. Xiang, Y. Wang, S.
AC

Zhang, J. Power Sources, 403 (2018) 137-156.

[23] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M. J. Rhodes, C.-Z. Li, Fuel,
88 (2009) 1810-1817

[24] M. Garcia-Perez, X.S. Wang, J. Shen, M.J. Rhodes, F.J. Tian, W.J. Lee, H. Wu,
C.-Z. Li, Ind. Eng. Chem. Res.,47 (2008) 1846-1854.

82
ACCEPTED MANUSCRIPT

[25] M. Garcia-Perez, X.S. Wang, J. Shen, M.J. Rhodes, F.J. Tian, W.J. Lee, H. Wu,
C.-Z. Li, Ind. Eng. Chem. Res.,47 (2008) 1846-1854.

[26] A.V. Bridgwater, D. Meier, D. Radlein,


Org. Geochem., 30 (1999) 1479-1493.

T
[27] W.R.W. Isahak, M.W.M. Hisham, M.A. Yarmo, T.Yun. Hin, Renew. Sust. Energy.

IP
Rev., 16 (2012) 5910-5923.

CR
[28] T. Milne, F. Agblevor, M. Davis, S. Deutch, D. Johnson, Developments in
Thermochemical Biomass Conversion 1 (1997) 409-424.

US
[29] M. Garcia-Perez, A. Chaala, H. Pakdel, D. Kretschmer, D. Rodrigue, C. Roy,
AN
Energy Fuels,20 (2006) 364–375.

[30] J. Meng, A. Moore, D. Tilotta, S. Kelley, S. Park, ACS Sustain. Chem. Eng., 2
M

(2014) 2011−2018.
ED

[31] A. Oasmaa, D. Meier, CPL Press U.K., 3 (2005) 19-60.


PT

[32] M. Ahmad, A.U. Rajapaksha, J.E. Lim, M. Zhang, N. Bolan, D. Mohan,


M.Vithanage, S.S .Lee, Y.S. Ok, Chemosphere, 99 (2014) 19-33.
CE

[33] D. Mohan, A. Sarswat, Y.S. Ok, C.U.P. Jr, Bioresour. Technol.,160 (2014) 191-
AC

202.

[34] Y. Chen, H. Yang, X. Wang, S. Zhang, H. Chen, Bioresour. Technol.,107 (2012)


411-418.

[35] Y. Zhao, D. Feng, Y. Zhang, Y. Huang, S. Sun, Fuel Process. Technol., 141
(2016) 54-60.

83
ACCEPTED MANUSCRIPT

[36] M. Asadullah, S. Zhang, C.-Z. Li, Fuel Process. Technol., 91 (2010) 877-881.

[37] J. Zhang, J. Liu, R. Liu, Bioresour. Technol., 176 (2015) 288-291.

[38] S. Valin, J. Cances, P. Castelli, S. Thiery, A. Dufour, G. Boissonnet, B. Spindler,


Fuel, 88 (2009) 834-842.

T
IP
[39] G. Hu, H.Huang, Y.Li, Energy Fuels, 23 (2009) 1748-1753.

CR
[40] M.N. Uddin, W.M.A.W. Daud, H.F. Abbas, RSC Adv.,4 (2014) 10467-10490.

US
[41] Q. Liu, S. Wang, Y. Zheng, Z. Luo, K. Cen, J. Anal. Appl. Pyrolysis,82 (2008)
170-177.
AN

[42] M. He, B. Xiao, S. Liu, Z. Hu, X. Guo, S. Luo, J. Anal. Appl. Pyrolysis, 87 (2010)
181–187.
M

[43] S. Dasappa, P.J. Paul, H.S. Mukunda, G.N.K. Rajan, S. Sridhar, H.V. Sridhar,
ED

Curr. Sci., 87 (2004) 908–916.


PT

[44] R.Z. Vigouroux, Pyrolysis of biomass, 2001: Royal Institute of Technology.


CE

[45] M.D.M. Hasan, X.S. Wang, D. Mourant, R. Gunawan, C. Yu, X. Hu, S.


Kadarwati, M. Gholizadeh, H. Wu, B. Li, L. Zhang, C.-Z. Li, Fuel Process. Technol.,
AC

167 (2017) 215-220.

[46] T. Kan, V. Strezov,T.J. Evans, Renew. Sust. Energy. Rev., 57(2016)1126–1140.

[47] A. Demirbas, G. Arin, Energ. Source, 24 (2002) 471–82.

84
ACCEPTED MANUSCRIPT

[48] M. Amutio, G. Lopez, R. R. Aguado, J. Bilbao, M. Olazar, Energy Fuels, 26


(2012) 1353–62.

[49] M.I. Jahirul, M.G. Rasul, A.A. Chowdhury, N. Ashwath, Energies, 5 (2012)
4952–5001.

[50] M.J. Antal, H.L. Friedman, F.E. Rogers, Combust. Sci. Technol., 21 (1980) 141–

T
52.

IP
CR
[51] A. G. W. Bradbury, Y. Sakai, F. Shalizadeh, J. Appl. Polym. Sci.,23 (1979) 3271-
3280.

US
[52] B. Hu, Q. Lu, Y. Wu, Z. Zhang, M. Cui, D. Liu, C. Dong, Y. Yang, , J. Anal. Appl.
Pyrolysis, 134 (2018) 183-194.
AN

[53] M.J. Antal, G. Varhegyi, Ind. Eng. Chem. Res.,34 (1995) 703-717.
M

[54] A.L. Brown, D.C. Dayton, J.W. Daily, Energy Fuels, 15 (2001) 1286-1294.
ED

[55] M. Wua, G. Varhegyib, Q. Zha,Thermochim. Acta, 496 (2009) 59-65.


PT

[56] Y.-C. Lin, J. Cho, G.A. Tompsett, P.R. Westmoreland, G.W. Huber, J. Phys.
Chem. C, 113 (2009) 20097–20107.
CE

[57] P.R. Patwardhan, D.L. Dalluge, B.H. Shanks, R.C. Brown, Bioresour. Technol.,
AC

102 (2011) 5265–5269.

[58] J.B. Wooten, J.I. Seeman, M.R. Hajaligol, Energy-Fuels, 18 (2004) 1-15.

[59] D. K. Shen, S. Gu, Cellul. Chem. Technol., 2010, 44(1-3), 79–87.

85
ACCEPTED MANUSCRIPT

[60] T.E. McGrath, W.G. Chan, M.R. Hajaligol, J. Anal. Appl. Pyrolysis, 66 (2003)
51–70.

[61] R. Alen, E. Kuoppala, P. Oesch, J. Anal. Appl. Pyrolysis, 36 (1996) 137–148.

[62] A. Demirbas, Energy Convers. Manage., 41 (2000) 633–646.

T
[63] A. Demirbas, J. Anal. Appl. Pyrolysis,72 (2004) 243–248.

IP
CR
[64] X. Zhou, W. Li, R. Mabon, L.J. Broadbelt, Energy Technol., 5 (2016) 52-79.

[65] R.S. Miller, J. Bellan, Combust. Sci. Technol., 126 (1997) 97-137.

US
[66] E. Ranzi, A. Cuoci, T. Faravelli, A. Frassoldati, G. Migliavacca, S. Pierucci, S.
AN
Sommariva, Energy-Fuels 22 (2008) 4292 – 4300.

[67] W. Shu-Rong, L. Tao, R.U. Bin, G. Xiu-Juan, Chem. J. Chinese U., 29 (2013)
M

782—787.
ED

[68] H. Yang, R. Yan, H. Chen, D.H. Lee, C. Zheng, Fuel, 86 (2007) 1781–1788.
PT

[69] Y. Peng, S. Wu, J. Anal. Appl. Pyrolysis, 88 (2010) 134–9.


CE

[70] D.K. Shen, S. Gu, K.H. Luo, S.R. Wang, M.X. Fang, Bioresour. Technol., 101
(2010) 6136–6146.
AC

[71] F. Monteil-Rivera, M. Phuong, M. Ye, A. Halasz, J. Hawari, Ind. Crops Prod., 41


(2013) 356–364.

[72] S. Wang, K. Wang, Q. LiuQ, Y. Gu, Z. Luo, K. Cen, T. Biotechnol. Adv., 27


(2009) 562–567.

86
ACCEPTED MANUSCRIPT

[73] E. Adler, Wood Sci. Technol., 11 (1977) 169–218.

[74] W. Mu, H. Ben, A. Ragauskas, Y. Deng, BioEnerg. Res., 6 (2013) 1183–1204.

[75] D. Ferdous, A.K. Dalai, S.K. Bej, R.W. Thring, Energy-Fuels, 16 (2002)1405–
1412.

T
[76] D. Ferdous, A.K. Dalai, S.K. Bej, R.W. Thring, N.N. Bakhshi NN, Fuel Process.

IP
Technol., 70(2001) 9–26.

CR
[77] H. Ben, A.J. Ragauskas, Green Chem., 14 (2012):72–76.

US
[78] R.K. Sharma, M.R. Hajaligol MR, J. Anal. Appl. Pyrolysis, 66 (2003) 123–144.
AN
[79] T. Hosoya, H. Kawamoto, S. Saka, J. Anal. Appl. Pyrolysis, 84 (2009) 79–83.

[80] S. Chu, A.V. Subrahmanyam, G.W. Huber, Green Chem., 15(2013) 125–136.
M

[81] C. Saiz-Jimenez, J.W. Deleeuw, Org. Geochem., 10(1986) 869–876.


ED

[82] G. Jiang, D.J. Nowakowski, A.V. Bridgwater, Energy-Fuels 24(2010) 4470–4475.


PT

[83] P.F. Greenwood, J.D.H. Heemst, E.A. Guthrie, P.G. Hatcher P.G, J. Anal. Appl.
CE

Pyrolysis, 62 (2002) 365–373.


AC

[84] H.E. Jegers, M.T. Klein, Ind. Eng. Chem. Process Des. Dev., 24 (1985) 173–183.

[85] R. Lou R, S.-B. Wu, G.-J. Lv, D.L. Guo, BioResources, 5 (2010) 2184–2194.

[86] Q. Liu, S. Wang, Y. Zheng, Z. Luo Z, K.Cen, J. Anal. Appl. Pyrolysis, 82 (2008)
170–177.

87
ACCEPTED MANUSCRIPT

[87] V.B.F. Custodis, P. Hemberger, Z. Ma, J. A.V. Bokhoven, J. Phys. Chem.,118


(2014) 8524–8531.

[88] X. Lin, S. Sui, S. Tan, C.U. Pittman, J. Sun, Z. Zhang, Energies, 2015, 8, 5107-
5121.

[89] D.J. Nowakowskia, A.V. Bridgwatera, D.C. Elliott, D. Meierc, P. Wild, J. Anal.

T
Appl. Pyrolysis, 88 (2010) 53–72.

IP
CR
[90] M. Brebu, G. Cazacu, O. Chirila, Cellul. Chem. Technol., 45 (2011) 43-50.

[91] R.P. Patwardhan, R.C. Brown, B.H. Shanks, ChemSusChem, 4 (2011) 1629 –
1636.
US
AN
[92] A. J. Caballero, J. A. Conesa, R. Font, A. Marcilla, J. Anal. Appl. Pyrolysis, 39
(1997) 161-183.
M

[93] C.E. Greenhalf, D.J. Nowakowski, A.B. Harms, J.O. Titiloye, A.V. Bridgwater,
Fuel, 108 (2013) 216-230.
ED

[94] T. Qu, W. Guo, L. Shen, J. Xiao, K. Zhao, Ind. Eng. Chem. Res., 50 (2011)
PT

10424–10433.
CE

[95] E. Butler, G. Devlin, D. Meier, K. McDonnell, Bioresour. Technol., 131 (2013)


202-209.
AC

[96] E. Biagini, F. Barontini, L. Tognotti, Ind. Eng. Chem. Res., 45 (2006) 4486-4493.

[97] T. Hosoya, H. Kawamoto, S. Saka, J. Anal. Appl. Pyrolysis, 78 (2007) 328-336.

[98] Q. Liu, Z. Zhong, S. Wang, Z. Luo, J. Anal. Appl. Pyrolysis,90 (2011) 213-218.

88
ACCEPTED MANUSCRIPT

[99] S. Wang, X. Guo, K. Wang, Z. Luo, J. Anal. Appl. Pyrolysis,91 (2011) 183-189.

[100] J. Zhang, Y.S. Choi, C.G. Yoo, T.H. Kim, R.C. Brown, B.H. Shanks, ACS
Sustain. Chem., 3 (2015) 293-301.

[101] F. Shafizadeh, P.P.S. Chin, ACS Symp. Ser., 43 (1977) 57–81.

T
[102] C. Branca, C.D. Blasi, J. Anal. Appl. Pyrolysis, 67 (2003) 207–19.

IP
CR
[103] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Fuel,
88 (2009) 1810–1817.

US
[104] S. Luo, B. Xiao, Z. Hu, S. Liu, Y. Guan, L. Cai, Bioresour. Technol.,101 (2010)
6517-6520.
AN

[105] H. Bennadji, K. Smith, M.J. Serapiglia, E.M. Fisher, Energy-Fuels 28 (2014)


7527–7537.
M

[106] S. Sensoz, D. Angin, S. Yorgun, Biomass Bioenergy, 19 (2000) 271–279.


ED

[107] H. Lu, E. Ip, J. Scott, P. Foster, M. Vickers, L.L. Baxter, Fuel, 89 (2010) 1156–
PT

1168.
CE

[108] N. Ali, M. Saleem, K. Shahzad, S. Hussain, A. Chughtai, Pol. J. Chem., 18


(2016) 88-96.
AC

[109] S. Septien, S. Valin, C. Dupont, M. Peyrot, S. Salvador, Fuel, 97 (2012) 202–


210.

[110] J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M.J. Rhodes, C.-Z. Li, Fuel,
88 (2009) 1810–1817.

89
ACCEPTED MANUSCRIPT

[111] D.V. Suriapparao, R. Vinu, Waste and Biomass Valori., 9 (2018) 465–477.

[112] C.K.W. Ndiema, P.N. Manga, C.R. Ruttoh, Energy Convers. Manage., 43
(2002) 2157–2161.

[113] Y. Li, H. Liu, Biomass Bioenergy, 19 (2000) 177–186.

T
[114] S. Mani, T.G. Tabil, S. Sokhansanj, Bioresour. Technol., 97 (2006) 1420–1426.

IP
CR
[115] L.G. Tabil, S. Sokhansanj, Powder Handling Process 8, (1996) 17–23.

[116] L.G. Tabil, S. Sokhansanj, Powder Handling Process 8, (1996) 117–122.

US
[117] K.V.S. Sastry, D.W. Fuerstenau, Powder Technol., 7 (1973) 97–105.
AN

[118] D. Carpenter, T.L. Westover, S. Czernika, W. Jablonski, Green Chem., 16


(2014) 384-406.
M

[119] C. Yu, P. Thy, L. Wang, S.N. Anderson, J.S. VanderGheynst, S.K. Upadhyaya,
ED

B.M. Jenkins, Fuel Process. Technol., 128 (2014) 43-53.


PT

[120] L. Deng, T. Zhang, D. Che, Fuel Process. Technol., 106 (2013) 712-720.
CE

[121] L. Jiang, S. Hu, L.-S. Sun, S. Su, K. Xu, L.-M. He, J. Xiang, Bioresour. Technol.,
146 (2013) 254-260.
AC

[122] R.C. Brown, ACS Symp. Ser.,784 (2001) 123-132.

[123] A.O. Ajani, S.E. Agarry, O.O. Agbede, J. App. Sci. Environ Manage. 15 (2013)
531-537.

90
ACCEPTED MANUSCRIPT

[124] M.R. Gray, W.H. Corcoran, G.R. Gavalas, Ind. Eng. Chem. Process Des. Dev.,
24 (1985) 646–651.

[125] B.C. Saha, B.L. Iten, M. Cotta, Y.V. Wu, Biotechnol. Progr., 21 (2005) 3816–
3822.

[126] Z. Zhu, R. Simister, S. Bird, S.J. McQueen-Mason, L.D. Gomez, D.J.

T
Macquarrie, AIMS Bioeng., 2 (2015) 449-468.

IP
CR
[127] H. Wang, R. Srinivasan, F. Yu, P. Steele, Q. Li, B. Mitchell, Energy-Fuels 25
(2011) 3758–3764.

US
[128] X. Wang, S. Leng, J. Bai, H. Zhou, X. Zhong, G. Zhuang, J. Wang, R. S. C., 5
(2015) 24984–24989.
AN

[129] J. Lee, C.J. Houtman, H.Y. Kim, I.G. Choi, T.W. Jeffries, Bioresour. Technol.,
102 (2011) 7451–7456.
M

[130] J. Lee, T.W. Jeffries, Bioresour. Technol., 102 (2011) 5884–5890.


ED

[131] T. Marzialetti, M.B.V. Olarte, C. Sievers, T.J.C. Hoskins, P.K. Agrawal, C.W.
PT

Jones, Ind. Eng. Chem. Res., 47 (2008) 7131–7140.


CE

[132] F. Carrillo, M.J. Lis, X. Colom, M. Lopez-Mesas, J. Valldeperas, Process


Biochem., 40 (2005) 3360-3364.
AC

[133] K.-H. Lee, B.-S. Kang, Y.-K. Park, J.-S. Kim, Energy Fuels, 19 (2005) 2179–
2184.

[134] N. Mosier, C. Wyman, B. Dale, R. Elander, Y.Y. Lee, M. Holtzapple, M. Ladisch,


Bioresour. Technol., 96 (2005) 673-686.

91
ACCEPTED MANUSCRIPT

[135] Y.S. Cheng, Y. Zheng, C.W. Yu, T.M. Dooley, B.M. Jenkins, J.S.V. Gheynst,
Appl. Biochem. Biotechnol., 162 (2010) 1768–1784.

[136] R. Sun, J.M. Lawther, W.B. Banks, Ind. Crops Prod.,2 (1995) 127–145.

[137] Md. Suruzzaman, ARPN J. Sci. Technol., 6 (2014) 354-358.

T
[138] Y. Zhao, Y. Wang, J.Y. Zhu, A. Ragauskas, Y. Deng Y, Biotechnol. Bioeng.,99

IP
(2008) 1320–1328.

CR
[139] S.K. Sharma, K.L. Kalra, H.S. Grewal, Biomass Bioenergy, 23 (2002) 237-243.
Environ. Prog. Sustain. Energy, 28 (2009) 435-440.

US
[141] S. Stephanidis, C. Nitsos, K. Kalogiannis, E.F. Lliopoulou, A.A. Lappas, K.S.
AN
Triantafyllidis, Catal. Today, 167 (2011) 37-45.

[142] Z. Liu, R. Balasubramanian, UAppl. Energy, 114 (2014) 857-864.


M

[143] A. Demirbas, J. Anal. Appl. Pyrolysis, 71 (2004) 803-815.


ED

[144] B.R. Crnogaca, E.F. Novi Materijali 26 (2017) 323-327.


PT

[145] K.J. Moscicki, L. Niedzwiecki, P. Owczarek, M. Wnukowski, J. Power Technol.,


CE

94 (2014) 233–249.
AC

[146] A. Zheng, Z. Zhao, S. Chang, Z. Huang, X. Wang, F. He, H. Li, Bioresour.


Technol., 128 (2013) 370-377.

[147] J. Meng, J. Park, D. Tilotta, S. Park, Bioresour. Technol., 111 (2012) 439-446.

[148] M.F. Li, L.X. Chen, X. Li, C.Z. Chen, Y.C. Lai, X. Xiao, Y.Y. Wu, Energy
Convers. Manage., 119 (2016) 463-472.

92
ACCEPTED MANUSCRIPT

[149] D. Chen, Y. Li, M. Deng, J. Wang, M. Chen, B. Yan, Q. Yuan Q, Bioresour.


Technol., 1 (2016) 615-622.

[150] D. Chen, J. Zhou, Q. Zhang, X. Zhu, Q. Lu, Bioresources, 9 (2014) 5893-5905.

[151] D. Chen, J. Zhou, Q. Zhang, Energy Fuels, 28 (2014) 5857−5863.

T
IP
[152] A.C. Louwes, L. Basile, R. Yukananto, J.C. Bhagwandas, E.A. Bramer, G.

CR
Brem, Biomass Bioenergy, 105 (2017) 116-126.

[153] S. Neupane, S. Adhikari, Z. Wang, A.J. Ragauskas, Y. Pu, Green Chem., 17


(2015) 2406–2417.
US
AN
[154] S. Ren, H. Lei, L. Wanga, Q. Bu, S. Chen, J. Wu, J. Julson, R. Ruan, Bioresour.
Technol., 135 (2013) 659–664.
M

[155] B. Ru, S. Wang, G. Dai, L. Zhang, Energy-Fuels, 29 (2015) 5865–5874.


ED

[156] A. Duque, P. Manzanares, I. Ballesteros, M. Ballesteros,Prog. Energy Combust.


Sci., 1 (2016) 349–368.
PT

[157] T. Pielhop, J. Amgarten, P.R.V. Rohr, M.H. Studer, Biotechnol Biofuels, 152
CE

(2016) 1-13.
AC

[158] P. Foody, Final Report to DOE, Contract AC02-79ET23050; Iotech: Ottawa,


Ontario, Canada, 1980.

[159] M. Wu, K. Chang, D. Gregg, A. Boussaid, R. Beatson, J. Saddler, Appl.


Biochem. Biotechnol., 47 (1999) 77−79.

93
ACCEPTED MANUSCRIPT

[160] S.M. Shevchenko, K. Chang, J. Robinson, J.N. Saddler, Bioresour. Technol.,


72 (2000) 207−211.

[161] K. Wang, J.X. Jiang, F. Xu, R.C. Sun, Bioresour. Technol., 100 (2009)
5288−5294.

[162] N. Jacquet, C. Vanderghem, S. Danthine, N. Quievy, C. Blecker, J. Devaux, M.

T
Paquot, Bioresour. Technol., 121 (2012) 221−227.

IP
CR
[163] Z. Sebestyen, E. Jakab, Z. May, B. Sipos, K. Reczey, J. Anal. Appl. Pyrolysis,
101 (2013) 61–71.

US
[164] R.H. Atalla, Gordon and Breach Science, New York (1988) 7119.
AN
[165] X.F. Sun, F. Xu, R.C. Sun, Z.C. Geng, P. Fowler, M.S. Baird, Carbohydr.
Polym., 60 2(005) 15−26.
M

[166] S. Jin, H. Chen, Biochem. Eng. J., 30 (2006) 225−230.


ED

[167] P.J. Morjanoff, P.P. Gray, Biotechnol. Bioeng., 29 (1987) 733−741.


PT

[168] Z. Chaula, M. Said, G. John, S. Manyele, C. Mhilu, Smart Grid Renewa.,


Energy 5 (2014) 1-7.
CE

[169] E. Viola, M. Cardinale, R. Santarcangelo, A. Villone, F. Zimbardi, Biomass


AC

Bioenergy, 32 (2008) 613−618.

[170] H.-Z. Chen, Z.-H. Liu, Biotechnol. J.,10 (2015) 1-20.

[171] M.J. Negro, P. Manzanares, J.M. Oliva, I. Ballesteros, M. Ballesteros, Biomass


Bioenergy, 25 (2003) 301−308.

94
ACCEPTED MANUSCRIPT

[172] I. Ballesteros, J.M. Oliva, A.A. Navarro, A. GonzSalez, J. Carrasco, M.


Ballesteros, Appl. Biochem. Biotechnol., 84 (2000) 97−110.

[173] R. Martin-Sampedro, E.A. Capanema, I. Hoeger, J.C. Villar, O.J. Rojas, J.


Agric. Food. Chem., 59 (2011) 8761–8769.

[174] T. Pielhop, J. Amgarten, M.H. Studer, P.R. Rohr, Biotechnol Biofuels,, 10

T
(2017) 1-13.

IP
CR
[175] S.P.S. Chundawat, B. Venkatesh, B.E. Dale, Biotechnol. Bioeng., 96 (2007)
219-231.

US
[176] V. Balan, B. Bals, S.P. Chundawat, D. Marshall, B.E. Dale, Methods Mol Biol.,
581 (2009) 61-77.
AN

[177] M.W. Lau, C. Gunawan, B.E. Dale, Biotechnol Biofuels,,2 (2009) 30-30.
M

[178] G. Biersbach, B. Rijal, S.W. Pryor, W.R. Gibbons, Appl. Biochem.


Biotechnol.,177 (2015) 1530-1540.
ED

[179] B. Bals, C. Rogers, M. Jin, V. Balan, B. Dale, Biotechnol Biofuels,,3 (2010) 1-


PT

11.
CE

[180] M.J. Lau, M.W Lau, C. Gunawan, B.E. Dale, Appl. Biochem. Biotechnol.,162
(2010) 1847–1857.
AC

[181] F. Teymouri, L. Laureano-Perez, H. Alizadeh, B.E. Dale, Bioresour. Technol.,


96 (2005) 2014–2018.

[182] H. Alizadeh, F. Teymouri, T.I. Gilbert, B.E. Dale, Appl. Biochem.


Biotechnol.,124 (2005) 1133-1141.

95
ACCEPTED MANUSCRIPT

[183] Y. Zhang, P. Chen, S. Liu, P. Peng, M. Min, Y. Cheng, E. Anderson, N. Zhou, L.


Fan, C. Liu, G. Chen, Y. Liu, H. Lei, B. Li, R. Bioresour. Technol.230 (2017) 143-151.

[184] X. Wang, H. Chen, K. Luo, J. Shao, H. Yang, Energy Fuels, 22 (2008) 67–74.

[185] D. Chen, D. Li, X. Zhu, Energ. Sources, 37 (2015) 1005-1011.

T
[186] B.C. Saha, N. Qureshi, G.J. Kennedy, M.A. Cotta, Int. Biodeterior. Biodegrad.,

IP
109 (2016) 29-35.

CR
[187] P. Kumar, D.M. Barrett, M.J. Delwiche, P. Stroeve, Ind. Eng. Chem. Res., 48
(2009) 3713-3729.

US
[188] Y. Zeng, X. Yang, H. Yu, X. Zhang, F. Ma, Agric. Food Chem., 59 (2011) 9965-
AN
9971.

[189] X.W. Yang, Y.L. Zeng, F.Y. Ma, X.Y. Zhang, H.B. Yu, Bioresour. Technol., 101
M

(2010) 5475–9.
ED

[190] Y.Q. Yu, Y.L. Zeng, J.E. Zuo, F.Y. Ma, X.W. Yang, X. Zhang, Y. Yang,
Bioresour. Technol., 134 (2013) 198–203.
PT

[190] H. Suhara, S. Kodama, I. Kamei, N. Maekawa, S. Meguro, Int. Biodeterior.


CE

Biodegrad.,75 (2012) 176–180.


AC

[192] W. Du, H. Yu, L. Song, J. Zhang, C. Weng, F. Ma, X. Zhang, X, Biotechnol.


Biofuels, 37 (2011) 1-8.

[193] M. Taha, E. Shahsavari, K. Al-Hothaly, A. Mouradov, A.T. Smith, A.S. Ball, E.M.
Adetutu, Appl. Biochem. Biotechnol., 175 (2015) 3709–3728.

96
ACCEPTED MANUSCRIPT

[194] T. Wang, C. Dong, Q. Lu, W. Jia, X. Yan, Z. Mao, ASABE Annual International
Meeting, Kansas City, Missouri (2013) 1-5.

[195] R. Castoldi, A. Bracht, G.R. Morais, M.L. Baesso, R.C.G. Correa, R.A. Peralta,
R.F.P.M. Moreira, M.T. Polizeli, C.G.M. Souz, R.M. Peralta, Chem. Eng. J., 258
(2014) 240–246.

T
[196] R. Potumarthi, R.R. Baadhe, P. Nayak, A. Jetty, Bioresour. Technol., 128

IP
(2013) 113–117.

CR
[197] L. Song, H. Yu, F. Ma, X. Zhang, BioResources, 8 (2013) 3802–3816.

US
[198] S. Cianchetta, B.D. Maggio, P.L. Burzi, S. Galletti, Appl. Biochem. Biotechnol.,
173 (2014) 609–623.
AN

[199] C. Wan, Y. Li, Bioresour. Technol., 102 (2011) 7507–7512.


M

[200] E. Rouchesa, M.-F. Dignac, H. Carrere, J. Anal. Appl. Pyrolysis, 123 (2017)
409-418.
ED

[201] R. Tapia-Tussell, J. Avila-Arias, J.D. Maldonado, D. Valero,E. Olguin-Maciel, D.


PT

Perez-Brito, L. Alzate-Gaviria, Energies, 11 (2018) 1-11.


CE

[202] K. Yan, F. Liu, Q. Chen, M. Ke, X. Huang, W. Hu, B. Zhou, X. Zhang, H. Yu,
Biotechnol Biofuels., 9 (2016) 1-11.
AC

[203] M. Saritha, A. Arora, Lata, Indian J. Microbiol., 52(2012) 122–130.

[204] E. Putun, F. Ates, A.E. Putun, Fuel, 87 (2008) 815-824.

[205] H. Zhang, R. Xiao, D. Wang, G. He, S. Shao, J. Zhang, Z. Zhong, Bioresour.


Technol., 102 (2011) 4258-4264.

97
ACCEPTED MANUSCRIPT

[206] C. Guizani, F.J. Escudero Sanz 1, S. Salvador, Fuel, 116 (2014) 310-320.

[207] Kwon Eilhann E, Jeon Young Jae, Yi Haakrho. Bioresour. Technol., 123 (2012)
673–677.

[208] Y.-F. Huang, W.-H. Kuan, C.-C. Chang, Y.-M. Tzou, Bioresour. Technol., 131

T
(2013) 274–280.

IP
CR
[209] T. Hanaoka, K. Sakanishi, Y. OkumuraFuel Process. Technol., 104 (2012)
287–294.

US
[210] E. Onal, B.B. Uzun, A.E. Putun, Int. J. Green Energy, 14 (2017) 1-8.
AN
[211] B. Zhang, J. Zhang, Energy Fuels, 31 (2017) 9627–9632.

[212] A.G. Borrego, L. Garavaglia, W.D. Kalkreuth, Int. J. of Coal Geol., 77 (2009)
M

409–15.
ED

[213] A. Demirbas, Energ. Sources, 29 (2007) 329–336.


PT

[214] F. Ates, M. A. Isikdag, Energy Fuels, 22 (2008) 1936–1943.


CE

[215] R.J.M. Westerhof, D.W.F. Brilman, W.P.M.V. Swaaij, S.R.A. Kersten, Ind. Eng.
Chem. Res.,49 (2010) 1160-1168.
AC

[216] P. Weerechanchai, C. Tanggsathitkulchai, M. Tanggsathitkulchai, Korean J.


Chem. Eng.,28 (2011) 2262-2272.

[217] R. Zanzi, K. Sjotrom, E. Bjornborn, Biomass Bio-energy, 23 (2002) 357-366.

[218] A.A. Rahman, N. Abdullah, F. Sulaiman, Adv. Energ. Eng., 2 (2014) 14-21.

98
ACCEPTED MANUSCRIPT

[219] A. Demirbas, Energ. Sources, 29 (2007) 329–336.

[220] N. Ibrahim, P.A. Jensen, K. Dam-Johansen, R.R. Ali, R.M. Kasmani Int. J.
Chem Mol. Eng., 6 (2012) 919-925.

[221] E.S. Noumi, J. Blin, J. Valette, P. Rousset, Energy Fuels, 29 (2015) 7301–7308.

T
IP
[222] M.R. Rovera, P.A. Johnstona, L.E. Whitmer, R.G. Smith, R.C. Brown, J. Anal.

CR
Appl. Pyrolysis, 105 (2014) 262–268.

[223] J. Sun, F. He, Y. Pan, Z. Zhang, ACTA Agr. Scand. B-S P 22 (2016) 1-11.

US
[224] G. Yang, Z. Wang, Q. Xian, F. Shen, C. Sun, Y. Zhang, J. Wu, RSC, 5 (2015)
AN
40117–40125.

[225] S. Zhou, M. Garcia-Perez, M.B. Pecha, S.R.A. Kersten, A.G. McDonald, R.J.M.
M

Westerhof, Energy Fuels, 27 (2013) 5867–5877.


ED

[226] P. Wild, H. Reith, E. Biofuels, 2 (2011)185-208.


PT

[227] M.L. Bras, J. Yvon, S. Bourbigot, V. Mamleev, J. Anal. Appl. Pyrolysis, 84


(2009) 1-17.
CE

[228] O. Onay, Fuel Process. Technol., 88 (2007) 523-531.


AC

[229] A. Bridgwater, Renew. Sust. Energy. Rev., 4 (2000) 1-73.

[230] E. Putun, B.B. Uzun, A.E. PutunBioresour. Technol., (2007) 701-710.

[231] F. Karaosmanoglu, E. Tetik, E. Gollu, Fuel Process. Technol.,59 (1999) 1-12.

99
ACCEPTED MANUSCRIPT

[232] D. Angin, Bioresour. Technol., 128 (2013) 593–597.

[233] H. Haykiri-Acma, S. Yaman, S. Kucukbayrak, Renew. Energ., 31 (2006) 803–


810.

[234] Q. Li, X. Fu, J. Li, Y. Wang, X. Lv, C. Hu, Energy Technol. 6 (2018) 366-378.

T
[235] B. Zhao, D. Connor, J. Zhang, T. Peng, Z. Shen, D.C.W. Tsang, D. Hou J.

IP
Clean. Prod., 174 (2018) 977-987.

CR
[236] C. Branca, D. Blasi, Ind. Eng. Chem. Res.,45 (2006) 5891-5899.

US
[237] F.L. Aquino, J.R. Hernandez, S.C. Capareda, An ASABE Meeting Presentation,
076083 (2007).
AN

[238] J. Gan, W. Yuan, W, An ASABE Meeting Presentation, 083719 (2008).


M

[239] J. Cai, Q. Wang, Q. Wang, Appl. Mech. Mater., 733 (2015) 280-283.
ED

[240] T.J. Morgan, S.Q. Turn, A.George, Plos One, 10 (2015) 1-28.
PT

[241] G. Newalkar, K. Iisa, A.D.D. Amico, C. Sievers, P. Agrawal, Energy Fuels,


2014, 28, 5144−5157.
CE

[242] D.S. Scott, J. Piskorz, D. Radlein, Ind. Eng. Chem. Process Des. Dev., 24
AC

(1985) 581-588.

[243] D.S. Scott, J. Piskorz, Can. J. Chem. Eng., 60 (1982) 666-674.

[244] D.S. Scott, R.L. Legge, J. Piskorz, P. Majerski, D. Radlein, Dev. in


thermochemi.l biomass conv., 1 (1997) 523-535.

100
ACCEPTED MANUSCRIPT

[245] A. Robson A, PyNe newsletter No. 11, UK: Aston University (2001) 1-2.

[246] R. McLellan, PyNe newsletter No. 10, UK: Aston University (2000) 12.

[247] C. Ming-Qiang, C. J. P. E., 6 (2006) 192-196.

[248] H. Seok, Y. Seok, S. Joon, Australia, Seventh Intern. Conf. CFD in the Miner.

T
Proc. Indu. (2009) 9-11.

IP
CR
[249] Q. Xiong, F. Xu, E. Ramirez, S. Pannala, S. Dawa, Fuel, 164 (2016) 11–17.

[250] K. Papadikis, S. Gu, A.V. Bridgwater, Energy Fuels, 2010, 24, 5634–5651.

US
[251] J.A. Garcia-Nunez, M.R. Pelaez-Samaniego, M.E. Garcia-Perez, I. Fonts, J.
AN
Abrego, R.J.M. Westerhof, M. Garcia-Perez, Energy Fuels, 31 (2017) 5751–5775.

[252] C. Rossi, R.G. Graham, UK: CPL Scientific Ltd, (1997) 300-306.
M

[253] S. Muller, Ensyn Technologies, PyNe newsletter, Bioenergy Research Group,


ED

UK: Aston University, 27 (2010) 11-12.


PT

[254] M.V. Velden, X. Fan, A. Ingram, J. Baeyens, The 12th Intern. Conf. Fluid.-New
Horiz. Fluid Eng, (2007) 897-904.
CE

[255] A.A. Lappas, M.C. Samolada, D.K. Iatridis, S.S. Voutetakis, I.A. Vasalos, Fuel,
AC

81 (2002) 2087–2095.

[256] A. Trendewicz, R. Braun, A. Dutta, J. Ziegler, Fuel, 133 (2014) 253–262.

[257] X. Dai, C. Wu, H. Li, Y. Chen, Energy Fuels, 14 (2000) 552-557.

[258] L. Junsheng, Intern. Conf. on Chal Environ Sci. Comp. Eng., (2010) 530-533.

101
ACCEPTED MANUSCRIPT

[259] B.M. Wagenaar, W. Prins, W.P.M. Swaaij, Chem. Eng. Sci., 24 (1994) 5109-
5126.

[260] J. Lede, J. Panagopoulos, H.Z. Li, J. Villermaux, Fuel, 64 (1985) 1514-1520.

[261] G.V.C. Peacocke, A.W. Bridgwater, Biomass Bioenergy, 7 (1995) 147-154.

T
IP
[262] G. Luo, D.S. Chandler, L.C.A. Anjos, R.J. Eng, P. Jia , F.L.P. Resende, Fuel,

CR
194 (2017) 229–238.

[263] B. Gomez-Monedero, F. Bimbela, J. Arauzo, J. Faria, M.P. Ruiz, Energy Fuels,


29 (2015) 1766−1775.
US
AN
[264] C.-Z. Li, X. S. Wang, H. Wu. PCT/AU2011/000741.2010.

[265] P. Brassard, S. Godbout, V. Raghavan, Biosyst. Eng., 161 (2017) 80-92.


M

[266] Y. Yu, Y. Yang, Z. Cheng, P.H. Blanco, R. Liu, A.V. Bridgwater, J. Cai, Energy
ED

Fuels, 30 (2016)10568-10574.
PT

[267] E. Le Roux, M. Chaouch, P.N. Diouf, T. Stevanovic, Biomass Bioenergy, 81


(2015) 202-209.
CE

[268] S. Liang, Y. Han, L. Wei, A.G. McDonald, Biomass Convers. Bio., 5 (2015)
AC

237-246.

[269] M. Garcia-Perez, T.T. Adams, J.W. Goodrum, D.P. Geller, K.C. Das, Energy
Fuels, 21(2007) 2363-2372.

[270] S.-S. Liaw, Z. Wang, P. Ndegwa, C. Frear, S. Ha, C.-Z. Li, C.-Z., M. Garcia-
Perez, J. Anal. Appl. Pyrolysis, 93 (2012) 52-62.

102
ACCEPTED MANUSCRIPT

[271] L. Ingram, D. Mohan, M. Bricka, P. Steele, D. Strobel, D. Crocker, B. Mitchell, J.


Mohammad, K. Cantrell, U. Charles, J. Pittman, Energy Fuels, 15 (2008) 614-625.

[272] N. Puy, R. Murillo, M.V. Navarro, M. V., Lopez, J. Rieradevall, G. Fowler, I.


Aranguren, G.T. Garcia, J. Bartroli, A.M. Mastral, Waste Manage., 31(2011) 1339-
1349.

T
IP
[273] P. Kim, A. Johnson, C.W. Edmunds, M. Radosevich, F. Vogt, T.G. Rials, N.

CR
Labbe, Energy Fuels, 25 (2011) 4693-4703.

[274] M.T. Morgano, H. Leibold, F. Richter, H. Seifert, J. Anal. Appl. Pyrolysis, 113
(2015) 216-224.
US
AN
[275] S.D. Ferreira, C.R. Altafini, D. Perondi, M. Godinho, Energy Convers. Manage.,
92 (2015) 223-233.
M

[276] R. Azargohar, K.L. Jacobson, E.E. Powell, A.K. Dalai, J. Anal. Appl. Pyrolysis,
104 (2013) 330-340.
ED

[277] J.N. Brown, R.C. Brown, Bioresour. Technol., 103 (2012) 405-414.
PT

[278] K. Raffelt, E. Henrich, A. Koegel, R. Stahl, J. Steinhardt, F. Weirich, Appl.


CE

Biochem. Biotechnol.,1 (2006) 153-164.


AC

[279] P. Brassard, G. Stephane, V. Raghavan, J.H. Palacios, M. Grenier, D. Zegan,


Energies, 10 (2017) 1-15.

[280] M. Mathew, L. Muruganandam, Int. J. Chem. React. Eng., 15 (2017) 1-12.

[281] S. Thangalazhy-Gopakumar, S. Adhikari , H. Ravindran, R.B. Gupta, O. Fasina,


M. Tu, S.D. Fernando, Bioresour. Technol., 101 (2010) 8389–8395.

103
ACCEPTED MANUSCRIPT

[282] N. Dewayanto, M.R. Nordin, Chemica, 2 (2015) 29-37.

[283] K. Smets, A. Roukaerts, J. Czech, G. Reggers, S. Schreurs, R. Carleer, J.


Yperman, Biomass Bioenergy, 57 (2013) 180–190.

[284] C. Torri, M. Reinikainen, C. Lindfors, D. Fabbri, A. Oasmaa, E. Kuoppala, J.

T
Anal. Appl. Pyrolysis,88 (2010) 7–13.

IP
CR
[285] T. Dickerson, J. Soria, Energies, 6 (2013) 514-538.

[286] C. Liu, H. Wang, A.M. Karim, J. Suna, Y. Wang, Catalytic fast pyrolysis of

US
lignocellulosic biomass, SC, 23 (2014) 6-54.
AN
[287] T.R. Carlson, J. Jae, Y.-C. Lin, G.A. Tompsett, G.W. Huber, J. Catal. 22 (2010)
110-124.
M

[288] J. Jae, G.A. Tompsett, A.J. Foster, K.D. Hammond, S.M. Auerbach, R.F. Lobo,
G.W. Huber , J. Catal. 279 (2011) 257-268.
ED

[289] Y. Zhao, L. Deng, B. Liao, Y. Fu, Q.-X. Guo, Energy Fuels 24 (2010) 5735-
PT

5740.
CE

[290] T.R. Carlson, Y.T. Cheng, J. Jae, G.W. Huber, Energ Environ Sci, 4 (2011)
145–161.
AC

[291] A. Zheng, L. Jiang, Z. Zhao, Z. Huang, K. Zhao, G. Wei, H. Li, WIREs Energy
and Environ., 6 (2017) 1-18.

[292] J.-Y. Kim, J.H. Lee, J. Park, J.K. Kim, D. An, I.K. Song, J.W. Choi, J. Anal. Appl.
Pyrolysis, 114 (2015) 273-280.

104
ACCEPTED MANUSCRIPT

[293] T. Ohra-Aho, J. Linnekoski, J. Anal. Appl. Pyrolysis, 113 (2015) 186-192.

[294] H.W. Lee, Y.-M. Kim, J. Jae, B.H. Sung, S.-C. Jung, S.C. Kim, J.-K. Jeon, Y.-K.
Park, J. Anal. Appl. Pyrolysis, 122 (2016) 282-288.

[295] X. Guo, S. Wang, Y. Zhou, Z. Luo, Intern. J. Energ. Environ., 5 (2011) 524-531.

T
[296] S. Wang, B. Ru, H. Lin, W. Sun, C. Yu, Z. Luo, Chem. J. Chinese U., 30 (2014)

IP
848–854.

CR
[297] T. Shun, Z. Zhijun, S. Jianping, W. Qingwen, Chin. J. Catal., 34 (2013) 641–
650.

US
[298] S. Wan, Y. Wang, Front. Chem. Scie Eng, 24 (2014) 1-15.
AN

[299] G. Yildiz, M. Pronk, M. Djokic, K.M. Geem K M, F. Ronsse, R. Duren, W. Prins,


J. Anal. Appl. Pyrolysis, 103 (2013) 343–351.
M

[300] T.S. Nguyen, M. Zabeti, L. Lefferts, G. Brem, K. Seshan,. Biomass Bioenergy,


ED

48 (2013) 100–110.
PT

[301] A. Gungor, S. Onenc, S. Ucar S, J. Yanik, J. Anal. Appl. Pyrolysis, 97 (2012)


39–48.
CE

[302] Michael Stocker, Renew. Resources, 2008, (2008) 9200 – 9211.


AC

[303] R. French, S. Czernik, Fuel Process. Technol., 91 (2010) 25-32.

[304] A.A. Lappas, K.G. Kalogiannis, E.F. Iliopoulou, K.S. Triantafyllidis, S.D.
Stefanidis, WIREs Energy and Environ., 1 (2012) 285-297.

[305] G. Kabir, B.H. Hameeda, Renew. Sust. Energy. Rev., 70 (2017) 945–967.

105
ACCEPTED MANUSCRIPT

[306] C.A. Mulle``n, A.A. Boateng. ACS Sustain. Chem. Eng., 3 (2015) 1623–1631.

[307] W.-L. Fanchiang, Y.-C. Lin, Appl. Catal. A, A, 419 (2012) 102–110.

[308] J. Fermoso, H. Hernando, P. Jana, I. Moreno, J. Prech, O. Hernandez, P.


Pizarro, J.M. Coronado, J. Cejka, D.P. Serrano, Catal. Today, 277 (2016) 171-181.

T
IP
[309] E.F. Iliopoulou, E.V. Antonakou, S.A. Karakoulia, I.A. Vasalos, A.A. Lappas,

CR
K.S. Triantafyllidis, Chem. Eng. J., 134 (2007) 51–57.

[310] J. Adam, E. Antonakou, A. Lappas, M. Stocker, M.H. Nilsen, A. Bouzga, J.E.

US
Hustad, G. OyeMicroporous Mesoporous Mater., 96 (2006) 93–101.
AN
[311] S.C. Du, J.A. Valla, G.M. Bollas, Green Chem., 15 (2013) 3214–3229.

[312] G. Duman, M. Pala, S. Ucar, J. Yanik, J. Anal. Appl. Pyrolysis, 103 (2013)
M

352–61.
ED

[313] A. Pattiya, J.O. Titiloye, A.V. Bridgwater, J. Anal. Appl. Pyrolysis, 88 (2008) 72-
79.
PT

[314] P.T. Williams, N. Nugranad, Energy, 25 (2000) 493-513.


CE

[315] P. Pan, C.W. Hu, W.Y. Yang, Y.S. Li, L.L. Dong, L.F. Zhu, D.M. Tong, R.W.
AC

Qing, Y. Fan, Bioresour. Technol., 101 (2010) 4593-4599.

[316] H. Zhang, R. Xiao, H. Huang, G. Xiao, Bioresour. Technol., 100 (2009) 1428-
1434.

[317] C.H. Engtrakul, C. Mukarakate, A.K. Starace, K.A. Magrini, A.K. Rogers, M.M.
Yung, Catal. Today, 269 92016) 175–181.

106
ACCEPTED MANUSCRIPT

[318] K. Giannakopoulou, M. Lukas, A. Vasiliev, C. Brunner, H. Schnitzer, Bioresour.


Technol., 101 (2010) 3209–3219.

[319] A. Aho, N. Kumar, K. Eranen, T. Salmi, M. Hupa, D.Y. Murzin, Fuel, 87 (2008)
2493-2501.

T
[320] J. Adam, M. Blazso, E. Meszaros, M. Stocker, M.H. Nilsen, A. Bouzga, J.E.

IP
Hustad, M. Gronli, G. Oye, Fuel, 84 (2005) 1494-1502.

CR
[321] J. Li, X. Li, G. Zhou, W. Wang, C. Wang, S. Komarneni, Y. Wang, Appl. Catal.
A, 470 (2014) 115–122.

US
[322] Y.T. Cheng, Z.P. Wang, C.J. Gilbert, W. Fan, G.W. Huber, Angew. Chem. Int.
AN
Ed.51 (2012) 11097–11100.

[323] T.S. Nguyen, S. He, G. Raman, K. Seshan, Chem. Eng. J.,229 (2016) 415–419.
M

[324] S. Karnjanakom, G. Guan, B. Asep, X. Du, X. Hao, J. Yang, C.H. Samart, A.


ED

Abudula, RSC Adv.,5 (2015), 83494–83503.


PT

[325] S. Karnjanakom, A. Bayu, P. Xiaoketi, X. Hao, S. Kongparakul, C.H. Samart, A.


Abudula, G. Guan, RSC Adv., 6 (2016), 50618–50629.
CE

[326] O.F. Mante, J.A. Rodriguez, S.D. Senanayake, S.P. Babu, Green Chem., 17
AC

(2015) 2362–2368.

[327] J. Grams, M. Niewiadomski, A.M. Ruppert, W. Kwapinski, C. R. Chim., 18


(2015) 1223–1228.

[328] J. Grams, A.M. Ruppert, Energies, 10 (2017) 900-924.

107
ACCEPTED MANUSCRIPT

[329] M.I. Nokkosmaki, A.O.I. Krause, E.A. Leppamaki, E.T. Kuoppala, , Catal.
Today, 4 (1998) 405-409.

[330] D. Rocha, C.A. Luengo, C.E. Snape, Org. Geochem., 30 (1999) 1527-1534.

[331] A.W. Williams, Georgia Institute of Technology, 2011.

T
[332] M.M. Barbooti, F.K. Matlub, H.M. Hadi,J. Anal. Appl. Pyrolysis, 98 (2012) 1–6.

IP
CR
[333] X. Zhang, L. Hanwu Lei, C. Shulin, J. Wu, Green Chem., 18 (2016) 4145-4169.

[334] C. Dorado, C.A. Mullen, A.A. Boateng, ACS Sustain. Chem. Eng., 2 (2014), 2
301–311.
US
AN
[335] F. Abnisa, W.M.A.W. Daud, Energy Convers. Manage., 87 (2014) 71–85.

[336] H.Y. Zhang, Y.T. Cheng, T.P. Vispute, R. Xiao, G.W. Huber, Energ. Environ
M

Sci., 4 (2011) 2297–2307.


ED

[337] H.Y. Zhang, T.R. Carlson, R. Xiao, G.W. Huber, Green Chem., 14 (2012) 98–
110.
PT

[338] T. Kuppens, T. Cornelissen, R. Carleer, J. Yperman, S. Schreurs, M. Jans M, T.


CE

J. Environ. Manage., 91 (2010) 2736–47.


AC

[339] J. Akhtar, N.S. Amin, J. Environ. Manage., 16 (2012) 5101–5109.

[340] I. Velghe, R. Carleer, J. Yperman, S. Schreurs, J. Anal. Appl. Pyrolysis, 92


(2011) 366–75.

[341] J.D. Martinez, A. Veses, A.M. Mastral, R. Murillo, M.V. Navarro, N. Puy, A.
Artigues, J. Bartroli, T. Garcia, Fuel Process. Technol., 119 (2014) 263–271.

108
ACCEPTED MANUSCRIPT

[342] M. Brebu, S. Ucar, C. Vasile, J. Yanik, Fuel, 89 (2010) 1911–1918.

[343] C.A. Mullen, A.A. Boateng, N.M. Goldberg, Energy Fuels, 27 (2013) 3867–
3874.

[344] C. Dorado, C.A. Mullen, A.A. Boateng, Ind. Eng. Chem. Res., 54 (2015) 9887–

T
9893.

IP
CR
[345] M. Ringer, V. Putsche, and J. Scahill, Technical Report, NREL/TP-510-37779
(2006).

US
[346] C. Jaroenkhasemmeesuk, N. Tippayawong, Energy Procedia 79 ( 2015 ) 950 –
955.
AN

[347] Z. Xiong, S.S.A. Syed-Hassan, J. Xu, Y. Wang, S. Hu, S. Su, S. Zhang, J.


Xiang, J. Anal. Appl. Pyrolysis, 134 (2018) 336-342.
M

[348] P.H. Moud, E. Kantarelis, K.J. Andersson, K. Engvall, Fuel, 211 (2018) 149–
ED

158.
PT

[349] M.Z. Stummann, M. Hoj, C.B. Schandel, A.B. Hansen, P. Wiwel, J. Gabrielsen,
P.A. Jensen, A.D. Jensen, Biomass Bioenergy, 115 (2018) 97–107.
CE

[350] C. Zhang, X. Hu, H. Guo, T. Wei, D. Dong, G. Hu, S. Hu, J. Xiang, Qi. Liu, Y.
AC

Wang, J. Anal. Appl. Pyrolysis, 134 (2018) 590-605.

109
ACCEPTED MANUSCRIPT

Table 1. The comparison of bio-oil produced from fast pyrolysis of wood and heavy
crude oil [3].
Properties Bio-oil Heavy fuel oil
Moisture content (wt%) 15.0–30.0 0.1
pH 2.5 Neutral
Specific gravity 1.20 0.94
C (wt%) 54–58 85

T
H (wt%) 5.5–7.0 11.0

IP
O (wt%) 35–40 1

CR
N (wt%) 0.0–0.2 0.3
S (wt%) trace 0.5–3.0
N (wt%)
Ash (wt%)
Higher heating value (MJ/kg)
US
trace
0.0-0.2
16–19
0.3
0.1
40
AN
Viscosity, at 500 °C (cP) 40–100 180
Solids (wt%) 0.2–1.0 1.0
M

Distillation residue (wt%) up to 50 1


Pour point (°C) –33 –18
ED

NOx emission (g/MJ) <0.7 N/A


SOx emission (g/MJ) 0 N/A
PT
CE
AC

110
ACCEPTED MANUSCRIPT

Table 2. Some of proposed analytical methods for bio-oil from wood pyrolysis [31].
Properties/analysis methods. Standards
Water content (wt %) ASTM 203
Conradson carbon residue content (wt %) ASTM D189
CHN content (wt %) ASTM D5291
Density, at 15 °C (kg/dm3) ASTM D4052

T
Viscosity, at 20 and 40 °C (cSt) ASTM D445

IP
Pour point (°C) ASTM D97

CR
Flash point (°C) ASTM D93

US
Table 3. The char yields and elemental analysis including surface area and

micropore volume for Eucalyptus pyrolysis at low heating rate (~10 °C/min) and high
AN
heating rate (>10 °C/min) at 800 °C [3].

Property Low heating rate High heating rate


M

Char yield (wt%, dry basis) 21.8 18.5


ED

C 88.23 81.85

H 0.71 1.55
PT

N 0.74 0.55

O 10.32 16.05
CE

Surface area (m2/g) 589 528

Micropore volume (mm3/g) 225 202


AC

111
ACCEPTED MANUSCRIPT

Table 4 Effect of temperature on the char physical properties (surface area and

micropore volume) for Eucalyptus pyrolysis at low heating rate (~10 °C/min) and high

heating rate (>10 °C/min) [3].

Property Low heating rate High heating rate

T
600 °C 700 °C 800 °C 900 °C 800 °C 900 °C

IP
Surface area 570 588 589 362 528 539

(m2/g)

CR
Micropore 218 224 225 138 202 206

volume (mm3/g)

US
AN
Table 5 Effect of pyrolysis temperature on the total carbon percentage in different
functional groups [37].
M

Temperature (°C) Total carbon (wt%)


Alkyl C O-alkyl C Aryl C Carboxylic C
ED

--- 10.29 66.13 20.02 3.56


200 10.89 53.83 29.37 5.91
400 11.68 6.47 75.79 6.06
PT

600 --- 12.80 81.85 5.35


CE
AC

112
ACCEPTED MANUSCRIPT

Table 6. The product distribution yields (on dry wood basis) from slow and fast
pyrolysis [4].
Type Conditions Solid (wt%) Liquid (wt%) Gas (wt%)
Slow ~400 °C, vapour residence time 35 30 35

T
hours→days

IP
Fast ~500 °C, vapour residence time ~1 12 75 13

CR
s

Cellulose
US
Table 7. Some of the biomass main components contents [2].
Biomass Hemicellulose Lignin Extractive Ash
AN
Pine 49.60 20.30 27.30 5.10 0.30
Poplar 49.00 24.00 20.00 5.90 1.00
Rice straw 37.00 16.50 13.60 13.10 19.80
M

Wheat 37.55 18.22 20.24 4.05 3.74


straw
ED

Corn straw 42.70 23.20 17.50 9.80 6.80


Bamboo 39.80 19.49 20.81 6.77 1.21
PT

Switchgrass 40.00– 31.00–35.00 6.00–12.00 5.00–11.00 5.00–6.00


45.00
CE
AC

113
ACCEPTED MANUSCRIPT

Table 8. Summary of process conditions of hemicellulose pyrolysis considering


global mechanism model [64].
Flow gas (rate, mL/min) Reactor Temperature and heating rate (°C, °C
min-1)

T
Nitrogen (150) TGA 110–500, 5

IP
Nitrogen (30) TGA 800, 5–20

CR
Nitrogen (60) TGA 25–800, 0.3
Nitrogen (60) TGA 25–800, 0.3

US
AN
M
ED
PT
CE
AC

114
ACCEPTED MANUSCRIPT

Table 9. The advantages and disadvantages of some of the reactors used in


biomass torrefaction [145].
Reactor type Advantages Disadvantages
Rotating drum - Uniform heat transfer - Low heat transfer
- Wide range of biomass size - Poor temperature control

T
- Various applications - Increased dust output

IP
- Limited upscaling ability
- High cost

CR
Screw - Relatively cheap - Uneven heating
- Lower chance of plugging - Poor heat exchange
- Proven technology - Limited scaling up potential
Herreshoff
oven
- Good heat transfer
US
- Good temperature control
- Wide range of biomass size
- Relatively large size of the
reactor
- Low sustainability
AN
- Scalable technology
Microwave - Uniform heat transfer - Low maturity
- Large specific throughput - Need for electricity
M

- Less dependent on biomass size - Dicult mass flow control


- Use of poor quality and
ED

non-uniform feedstock
- Good temperature
control
PT

Moving bed - simple and low cost - limited biomass size


- No moving parts inside - High pressure drop
CE

- Non-uniform temperature
- Difficult control
- Risk of unequal degree of
AC

devolatization and
decarbonisation
- Unproven scale up potential
Belt - Proven technology
- Excellent control
- Perfect plug flow
- various range of biomass sizes
- Relatively low cast

115
ACCEPTED MANUSCRIPT

Table 10. Summary of advantages and disadvantage of different biomass


pretreatment methods.
Method Advantages Disadvantages
Grinding - Decrease of cellulose crystallinity and - High power and energy

T
degree of polymerisation consumption

IP
- Reduction of particle size to increase
specific surface area and pore size

CR
Densification - Easy to transport and storage - High cost
- reduced cost of transportation due to
increased energy density

Concentrated acid
- reduced dust production
- High glucose yield
US - High cost of acid and need to
AN
- Ambient temperatures be recovered
- Corrosion-resistant
equipments are required
M

- Concentrated acids are toxic


and hazardous
Diluted acid - High recovery of sugars at the end of - Concentration of reducing
ED

the process sugars is relatively low


- Low formation of toxic products - Generation of degradation
products
PT

Alkali - Decrease in the degree of - High cost


polymerization and crystallinity of - Not used for large-scale plant
CE

cellulose
- Disruption of lignin structure
Torrefaction - Good decomposition of hemicellulose - Torrefied biomass carries too
AC

- Partial decomposition of cellulose and much fine particles during the


lignin feeding to pyrolysis reactor
- High carbon and energy content of
biomass
- Easy transport
- Low electricity consumption during
milling

116
ACCEPTED MANUSCRIPT

- Easy to store and resistant in front of


micro organism attack
Steam explosion - Causes lignin transformation and - Generation of toxic
hemicellulose solubilisation compounds
- Lower cost - Partial hemicellulose
- Higher yield of glucose and degradation
hemicellulose in the two-step method
Ammonia - Increases accessible surface area - Not very effective for the

T
- Less inhibitors formation biomass with high lignin content
fiberexpansion

IP
- Does not require small particle size of - High cost of large amount of
biomass ammonia

CR
Drying - improved density and - An energy-intensive process
durability of the fuel produced after
pyrolysis

US
- Low moisture content making the
biomass easier to compact
- Improve the efficiency of the pyrolysis
AN
- Increased bio-oil yield
Biological - Low energy requirements - Slow process rate
- Delignification - Very low treatment rate
M

- Reduction in degree of polymerization - Not very effective for


of cellulose commercial application
ED

- Partial hydrolysis of hemicelluloses


- No chemical requirements
- Mild environmental conditions
PT
CE
AC

117
ACCEPTED MANUSCRIPT

Table 11 Overview of fast pyrolysis reactor characteristics for bio-oil production [3,4].
Property Status Bio-oil Complexity Feed Inert gas Specific Scale Gas
yield on size specification requirements reactor up quality
dry size
biomass
(wt%)
Fluidised Commercial 75 Medium High High Medium Easy Low

T
bed
CFB & Commercial 75 High High High Medium Easy Low
Transported

IP
bed
Rotating Demonstration 70 High High Low Low Medium High
cone
Entrained Laboratory 60 Medium High High Medium Easy Low

CR
flow
Ablative Laboratory 75 High Low Low Low Difficult High
Screw or Pilot 60 Medium Medium Low Low Medium High
Auger
Vacuum None 60 High Low Low High Difficult Medium

US
Note: Commercial, demonstration, pilot and laboratory scale feedstock flow rates are

2 –20 ton/h, 200–2000 kg/h, 20–200 kg/h and 1–20 kg/h, respectfully.
AN
M
ED
PT
CE
AC

118
ACCEPTED MANUSCRIPT

Table 12 Summary of advantages and disadvantage of different reactors for


biomass pyrolysis [4,251].
Type Advantages Disadvantages
Bubbling fluidised - High bio-oil yield - Limit of biomass particle size
- Even temperature distribution - The risk of ash fusion
bed
eliminates hot spots.

T
- Catalyst is easily replaced or

IP
regenerated.
- More efficient contacting of gas and

CR
solid than in other catalytic reactors.
Circulating - High throughput - Increased reactor vessel size
- Uniform particle mixing - Pumping requirements and
fluidised bed and
transported bed
- Ability
continuous state
to US
- Uniform temperature gradients
operate reactor in
pressure drop
- Particle entrainment
- Lack of current understanding
AN
- Erosion of internal
components
- Pressure loss scenarios
M

Rotating cone - High efficiency of heat transferring - High energy consumption


- No inert gas needed.
ED

Ablative pyrolysis - Large particle can be used - Reaction rates limited by heat
- Inert gas is not required transfer to the reactor
- Controllable residence time - High cost to scale up
PT

- System is more intensive - High gas flow and production


- Good heat transfer
dilution
CE

Grinding pyrolysis - Wide range of biomass particle size - Plugging risk


- Easy to scale up - low organic yield
- High water content in bio-oil
AC

Auger - Low pyrolysis temperature - Plugging risk


- Compact, flexible design - Lower bio-oil yield
- No carrier gas, dilution - Moving parts in the hot zone
- High quality of char produced - Heat transfer limitation

119
ACCEPTED MANUSCRIPT

Table 13 The yields (wt%) of main products from the pyrolysis of rapeseed cake

using catalyst and non-catalyst [284].

Product No catalyst ɣ-Al2O3 HZSM-5 Na2CO3

Bio-oil 47.1 40.8 44.9 41.8

T
Char 23.0 22.6 23.1 19.0

IP
Gas 23.8 29.1 25.6 33.3

CR
US
AN
M
ED
PT
CE
AC

120
ACCEPTED MANUSCRIPT

Table 14 Some Comparisons of bio-oil yield between non-catalytic and catalytic

pyrolysis [282,283].

Catalyst Feedstock Yield without Yield with catalyst

catalyst (wt%) (wt%)

T
ZnO Pine sawdust 47.0 47.0

IP
SnO2 Pine sawdust 47.0 32.0

CR
SiO2 Pine sawdust 47.0 19.0

MgO Pine sawdust 44.0 16.0

CuO Pine sawdust US


47.0 49.0
AN
CaO/Al2O3 Pine sawdust 47.0 13.0

CaO/Al2O3 White pine powder 39.4 34.1


M

CaO/Al2O3 Sugarcane bagasse 23.0 24.0

CuO Sugarcane bagasse 23.0 34.0


ED

La2O3/ MgO Sugarcane bagasse 23.0 18.0

NiO Sugarcane bagasse 23.0 32.0


PT

CaO/Al2O3 Rapeseed cake 47.1 40.8


CE
AC

121
ACCEPTED MANUSCRIPT

Table 15 Aromatics selectivity (mol%-carbon) of bio-oil produced from different

feedstocks through catalytic pyrolysis using ZSM-5 catalyst [291].

Aromatics selectivity (mol%-carbon)

Cellulose Hemicellulose Lignin Pine Corncob Straw

T
Benzene 14.3 13.4 12.1 11.6 11.1 11.9

IP
Toulene 11.9 18.2 25.6 14.5 14.3 15.8

CR
Xylene 10.1 16.8 27.2 14.2 14.5 15.6

Benzene, 36.3 48.4 65.0 40.3 39.9 43.2

Toluene,

isomers
xylene US
AN

Ethylbenzene 2.6 3.3 1.7 2.5 3.1 3.0


M

Alkylbenzenes 5.7 8.0 7.8 8.4 7.4 7.9

Phenols 3.7 --- 1.4 3.7 2.8 2.5


ED

Indenes 13.4 9.0 3.0 10.7 12.6 11.6

Naphthalenes 25.9 24.0 17.7 24.8 25.2 24.1


PT

Benzofurans 2.0 1.2 --- 1.5 1.5 0.7


CE

Fluorenes, 9.2 2.7 3.4 6.9 5.7 5.4

Anthracenes and
AC

phenanthrenes

122
ACCEPTED MANUSCRIPT

Table 16 Summary of catalytic pyrolysis of biomass using zeolite catalysts under

different conditions [64].

Feedstock Catalyst Temperature Reactor type Process type

(°C)

Corn stalks ZSM-5, HY, USY 400–600 Tubular fixed- In situ and in

T
bed bed

IP
Pine wood Beta, Y, ZSM-5 and 450 Fluidised bed In situ

CR
Mordenite

Oak, corn cob, Ferrierite, Mordenite, 550 Fluidised bed In bed

corn stover,

switchgrass
Y, ZSM‐5, Beta
US
AN
Herb residue ZSM-5, Al-SBA-15 350–550 Fixed bed In bed

and alumina
M

Rice husk ZSM‐5 400–600 Fluidised bed In situ

Green HZSM‐5 300–500 Fixed bed In situ


ED

microalgae

Corncob HZSM-5 550 Fluidised bed In situ


PT

Olive Residue Clinoptilolite, ZSM-5, 350–500 Fixed bed In situ


CE

HY

Beech wood FCC, ZSM-5, MgO, 500 Fixed bed In situ


AC

Al2O3NiO, CrO2/TiO2

Beech wood HZSM-5, Silicalite, Al- 500 Fixed bed In situ

MCM-41

123
ACCEPTED MANUSCRIPT

Table 17 Summary of catalytic pyrolysis of biomass using non-zeolite catalysts

under different conditions [328].

Feedstock Catalyst Temperature Reactor type

T
(°C)

IP
Pine Na2CO3/Al2O3 500 Annular flow

CR
and Pt/Na2CO3/Al2O3

cedar Cu, Fe and Zn 500 Fixed bed

supported on

Al2O3 and SiO2


US
AN

Sunflower Cu and Fe 565 Fixed bed


M

stalks supported on

mesoporous
ED

rod-like Al2O3

Sugar maple TiO2, CeO2, 550 Py-GCMS


PT

CeOx-TiO2,
CE

ZrO2 and MgO

Wheat bran Pt/C, Pd/C, 700 Thermogravimetric


AC

ZSM-5, analyser

MCM-41

Pine sawdust Commercial 500–700 Conical spouted

Ni/Al2O3 bed

124
ACCEPTED MANUSCRIPT

Wood sawdust NiZnAlOx 535 Fixed bed

Beech, birch, Ni/CaO-ZrO2 500 Stirred batch

poplar and pine

wheat straw, Ni-Co/Mg-Al, Ni supported 400-650 Batch

timothy grass, on Al2O3, activated carbon,

T
canola meal TiO2, ZrO2, MgO,

IP
pine wood, Ni/biomass

CR
fruit pulp Ru/C

US
AN
M
ED
PT
CE
AC

125
ACCEPTED MANUSCRIPT

Table 18 Summary of the studies on the influence of co-pyrolysis of biomass with

plastics and tire waste [335].

Biomass Plastics/tier Temperat Ratio (wt/wt) Bio-oil yield (wt%) - Bio-oil yield

T
ure (°C) (biomass/plas biomass (wt%) - biomass

IP
tics or tier) and waste

Palm PS 500 1/1 46.13 61.63

CR
shell

Pine

cone
LDPE/PP/

PS
500 1/1
US 47.5 63.9/64.1/69.7
AN
Willow PHB 450 1/1 49.71 64.24

Willow PLA 450 1/1 48.85 51.30


M

Potato HDPE 500 1/1 23.00 39.00


ED

skin

Fir Electrical 500 1/1 46.30 62.30


PT

sawdus plastic

t waste
CE

Wood Block 500 1/1 39.30 63.10


AC

chip polypropyl

ene

Pine PE, 17% 400 1/1 32.00 53.00

residue PS and

27% PP

126
ACCEPTED MANUSCRIPT

Pinewo PS 450 1/1 46.00 67.00

od

sawdus

Willow Biopearls/ 450 1/1 50.10 52.79/59.24/5

T
Solanyl/ 1.52

IP
Potato

CR
starch

Wood Tier 500 1/1 42.80 8.30

Wood Tier 500


US
1/0-6/4-4/6-

0/1
45.00/46.20/47.00/

47.20
47.20
AN
\Wood Tier 500 1/0-9/1-8/2 47.60-50.00 47.60-
M

50.00

Note: PP (polypropylene), PE (polyethylene), LDPE (low density polyethylene), PS


ED

(polystyrene), PLA (polylactic acid), PHB (Polyhydroxybutyrate).


PT
CE
AC

127
ACCEPTED MANUSCRIPT

Table 19 Summary of the commercial or demonstration size pyrolysis reactors

[4,26,229,345].

Year Country Company Feeding rate Reactor type

T
(kg/h)

IP
1996 Canada Red Arrows - Ensyn 1667 Fluidised bed/riser

CR
2000 Canada Pyrovac 3500 Vacuum stirred

bed

2005

2013
Malaysia

Finland
Genting

Fortum - VALMET
US 2000

10000
Rotating cone

Fluidised bed/riser
AN
2014 Netherlands BTG-BTL/EMPYRO 5000 Rotating cone
M

2014 USA KiOR 21000 Catalytic pyrolysis

2015 Finland UPM 11574


ED

2017 Canada AE Cote-Nord Bioenergy 9000 Fluidised bed/riser

/ Ensyn
PT
CE
AC

128
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 1. Chemical structure of cellulose [13].


ED
PT
CE
AC

129
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 2. Yan-Fen et al. modified model of Bradbury for low temperature cellulose

pyrolysis [52].
AN
M
ED
PT
CE
AC

130
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 3. Wooten et al. offered model for cellulose pyrolysis [58].
AN
M
ED
PT
CE
AC

131
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 4. The proposed kinetic scheme for the cellulose pyrolysis by Shen et al. [59].
AN
M
ED
PT
CE
AC

132
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 5. Chemical decomposition pathway of cellulose in pyrolysis [53].


ED
PT
CE
AC

133
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Fig. 6. Chemical structure of hemicellulose [13].


PT
CE
AC

134
ACCEPTED MANUSCRIPT

T
IP
CR
Fig. 7. The one stage global mechanism model proposed for the pyrolysis of

hemicellulose [64].

US
AN
M
ED
PT
CE
AC

135
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 8. The multi stage global mechanism model proposed for the pyrolysis of
AN
hemicellulose by Miller et al. [65].
M
ED
PT
CE
AC

136
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED
PT

Fig. 9. The multi stage global mechanism model proposed for the pyrolysis of
CE

hemicellulose by Ranzi et al. [66].


AC

137
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 10. Chemical structure of xylan [67].
AN
M
ED
PT
CE
AC

138
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 11. Chemical decomposition pathway of xylan as a model compound of
AN
hemicellulose in pyrolysis [53].
M
ED
PT
CE
AC

139
ACCEPTED MANUSCRIPT

T
IP
CR
Fig. 12. The proposed mechanism for the xylan pyrolysis by Shu-Rong et al. [67].

US
AN
M
ED
PT
CE
AC

140
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Fig. 13. Chemical structure of lignin [13].


PT
CE
AC

141
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
Fig. 14. The proposed two pathway decomposition of lignin during the pyrolysis [92].
M
ED
PT
CE
AC

142
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 15. Proposed lignin decomposition mechanism during the pyrolysis [74].
AN
M
ED
PT
CE
AC

143
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN

Fig. 16. Summarised biomass pyrolysis mechanism [13].


M
ED
PT
CE
AC

144
ACCEPTED MANUSCRIPT

T
Fig. 17. One-component mechanism for biomass pyrolysis [101].

IP
CR
US
AN
M
ED
PT
CE
AC

145
ACCEPTED MANUSCRIPT

T
IP
CR
Fig. 18. Multi-component kinetics for biomass pyrolysis [102].

US
AN
M
ED
PT
CE
AC

146
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 19. A schematic of bio-oil production by fast pyrolysis using torrefaction as a


ED

pretreatment process [152].


PT
CE
AC

147
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED
PT
CE
AC

Fig. 20. A schematic of ammonia fibre expansion reactor [176].

148
ACCEPTED MANUSCRIPT

T
IP
CR
Fig. 21. Different heating methods in pyrolysis reactors.

US
AN
M
ED
PT
CE
AC

149
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 22. A common schematic of bubbling fluidised bed reactor [4].
AN
M
ED
PT
CE
AC

150
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
Fig. 23. A common schematic of circulating fluidised bed reactor [4].
M
ED
PT
CE
AC

151
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 24. A schematic of rotating cone reactor for pyrolysis of biomass [4].
AN
M
ED
PT
CE
AC

152
ACCEPTED MANUSCRIPT

T
IP
Fig. 25. A schematic of ablative reactor used for pyrolysis of biomass [4].

CR
US
AN
M
ED
PT
CE
AC

153
ACCEPTED MANUSCRIPT

T
IP
CR
Fig. 26. Grinding pyrolysis reactor schematic [45].

US
AN
M
ED
PT
CE
AC

154
ACCEPTED MANUSCRIPT

T
Fig. 27. Auger reactor schematic for biomass pyrolysis [265].

IP
CR
US
AN
M
ED
PT
CE
AC

155
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
Fig. 28. Chemical composition of bio-oil produced from the pyrolysis of wood [286].
M
ED
PT
CE
AC

156
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 29. Reaction pathway of catalytic pyrolysis of cellulose [291].
AN
M
ED
PT
CE
AC

157
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED
PT
CE
AC

Fig. 30. Reaction pathway of catalytic pyrolysis of hemicellulose and lignin [291].

158
ACCEPTED MANUSCRIPT

T
IP
CR
US
Fig. 31. Reaction pathway of catalytic pyrolysis of biomass [297].
AN
M
ED
PT
CE
AC

159
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Fig. 32. The functional groups in the bio-oil produced from the pyrolysis of rapeseed
PT

cake: (a) organic phase of bio-oil (b) aqueous phase of bio-oil [282].
CE
AC

160
ACCEPTED MANUSCRIPT

T
Fig. 33. Reaction pathway of catalytic pyrolysis of hollocellulose biomass over
ZSM-5 catalyst [305].

IP
The graphical abstract need to be provided. For the graphical abstract, there

CR
should be a description portion containing 20-35 words.

US
AN
Graphical Abstract
M
ED
PT
CE

Description
AC

Pyrolysis is a promising process to convert the biomass into bio-oil. This study
reviewed the progress in pyrolysis in terms of the reaction, process, pyrolysers, main
parameters and the status of commercialisation of pyrolysis.

161

Das könnte Ihnen auch gefallen