Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Bioresource Technology 342 (2021) 126067

Contents lists available at ScienceDirect

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

Fast hydropyrolysis of biomass Conversion: A comparative review


Shinyoung Oh a, 1, Jechan Lee b, 1, Su Shiung Lam c, 1, Eilhann E. Kwon d, 1, Jeong-Myeong Ha a, 1,
Daniel C.W. Tsang e, Yong Sik Ok f, Wei-Hsin Chen g, h, i, Young-Kwon Park j, *
a
Clean Energy Research Center, Korea Institute of Science and Technology, Seoul 02792, Republic of Korea
b
Department of Environmental and Safety Engineering & Department of Energy Systems Research, Ajou University, Suwon 16499, Republic of Korea
c
Pyrolysis Technology Research Group, Higher Institution Centre of Excellence (HICoE), Institute of Tropical Aquaculture and Fisheries (AKUATROP), Universiti
Malaysia Terengganu, 21030 Kuala Nerus, Terengganu, Malaysia
d
Department of Environment and Energy, Sejong University, Seoul 05006, Republic of Korea
e
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong, PR China
f
Korea Biochar Research Centre, APRU Sustainable Waste Management Program & Division of Environmental Science and Ecological Engineering, Korea University,
Seoul 02841, Republic of Korea
g
Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 701, Taiwan
h
Research Center for Smart Sustainable Circular Economy, Tunghai University, Taichung 407, Taiwan
i
Department of Mechanical Engineering, National Chin-Yi University of Technology, Taichung 411, Taiwan
j
School of Environmental Engineering, University of Seoul, Seoul 02504, Republic of Korea

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

• Studies on hydropyrolysis and catalytic


hydropyrolysis were critically reviewed.
• The pyrolysis/hydropyrolysis and non-
catalytic/catalytic processes were
compared.
• Integrated hydrotreating processes were
developed to enhance hydrocarbon
yield.
• Combined hydropyrolysis process could
economically produce amounts of
hydrocarbons.

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

Keywords: Recent studies show that fast hydropyrolysis (i.e., pyrolysis under hydrogen atmosphere operating at a rapid
Hydropyrolysis heating rate) is a promising technology for the conversion of biomass into liquid fuels (e.g., bio-oil and C4+
Pyrolysis hydrocarbons). This pyrolysis approach is reported to be more effective than conventional fast pyrolysis in
Biofuels
producing aromatic hydrocarbons and also lowering the oxygen content of the bio-oil obtained compared to
Hydrocarbons
Biorefinery
hydrodeoxygenation (a common bio-oil upgrading method). Based on current literature, various non-catalytic
and catalytic fast hydropyrolysis processes are reviewed and discussed. Efforts to combine fast hydropyrolysis
and hydrotreatment process are also highlighted. Points to be considered for future research into fast hydro­
pyrolysis and pending challenges are also discussed.

* Corresponding author.
E-mail address: catalica@uos.ac.kr (Y.-K. Park).
1
These authors are co-first authors because they contributed equally to this work.

https://doi.org/10.1016/j.biortech.2021.126067
Received 26 August 2021; Received in revised form 27 September 2021; Accepted 29 September 2021
Available online 1 October 2021
0960-8524/© 2021 Elsevier Ltd. All rights reserved.
S. Oh et al. Bioresource Technology 342 (2021) 126067

1. Introduction with thermal decomposition of biomass. The presence of hydrogen in


FHP, however, results in the addition of complex hydrotreating re­
Liquid fuels comprise hydrocarbon mixtures with the advantages of actions that occur simultaneously or successively (Stummann et al.,
high-energy density and ease of use. The production of high-value liquid 2020a). Such reactions can involve C–C coupling, hydrocracking,
fuels from non-fossil resources to reduce fossil fuel consumption has alkylation, decarboxylation, decarbonylation, hydrogenation, hydro­
become an imperative need. Biomass is a sustainable, renewable, and deoxygenation (HDO), and recombination (Stummann et al., 2021).
near carbon–neutral source that has the potential to substitute for pe­ Fig. 1 describes reaction chemistries and mechanisms for conventional
troleum via hydrocarbon fuel production (Ahmad et al., 2021). The fast pyrolysis, non-catalytic fast hydropyrolysis (NCFHP), and CFHP of
production of hydrocarbon-based liquid fuels or chemicals from biomass biomass samples (He et al., 2020). The effect of HDO caused by hy­
(including waste biomass) has been studied widely, as the utilization of drocracking and deoxygenation is more pronounced in NCFHP and
renewable hydrocarbons obtained from biomass conversion could CFHP than in conventional fast pyrolysis. The use of a bifunctional
circulate the carbon and this reduces global warming (Jung et al., 2020). catalyst (e.g., Rh/ZrO2) increases alkanes and decreases oxygenates,
Among various biomass conversion processes, thermochemical conver­ indicating that CFHP is more effective at removing oxygen from pyrol­
sion shows potential for commercialization (Shahabuddin et al., 2020). ysis bio-oil and improving bio-oil quality than NCFHP.
Pyrolysis is a thermochemical conversion process that can be per­ C–C coupling reactions lead to the production of light compounds
formed with various reactor types, inert gas types, residence times, and present in both gaseous and liquid phase but cause polymerization and
heating rates (Jung et al., 2018). This process is applicable to the coke formation on the catalyst. Hydrocracking reactions reduce the
treatment of various types of biomass (Moogi et al., 2020), organic waste carbon chain length of pyrolytic products to a length within the range
(Kwon et al., 2018), municipal solid waste (Lee et al., 2020a), plastics C4–C21, representing the range comparable to gasoline, jet-fuel, and
(Hassan et al., 2019), and industrial waste (Chen et al., 2020a). Among diesel-range hydrocarbons. However, they also facilitate the formation
different pyrolysis processes, fast pyrolysis is widely employed to pro­ of light non-condensable C1–C3 hydrocarbons, which deteriorates
duce liquid fuels from biomass (e.g., bio-oil). Fast pyrolysis typically overall hydrocarbon yield in bio-oil.
operates at moderate temperatures (400–600 ◦ C), rapid heating rate Acid catalyzed alkylation is favorable to occur at temperatures at
(>100 ◦ C min− 1), and short residence time of pyrolytic vapors (0.5–2 s) which FHP is conducted. This reaction can enhance carbon recovery in
(Lee et al., 2019). High contents of oxygen in bio-oil leads to thermal condensable organic phase by incorporating short hydrocarbons into
instability and corrosiveness of bio-oil and lowers the heating value aromatic compounds (Resasco et al., 2016). Alkylation reactions also
compared to petroleum-derived oil (Lee et al., 2020b). To tackle the occur via alcohols or olefins generated during the FHP (Resasco et al.,
problems of bio-oil, it needs to be upgraded properly to improve its 2016).
quality. One bio-oil upgrading route involves passing the vapor from the Decarboxylation and decarbonylation reactions also occur during
pyrolysis process through a catalyst bed prior to condensing the vapor to FHP and in turn decrease the oxygen content in the condensable com­
form bio-oil (Mo et al., 2020). Typically, zeolites are used in this type of pounds by removing CO2 and CO. These reactions help stabilizing the
upgrading process (referred as catalytic fast pyrolysis) (Kumar et al., reactive oxygenated phase in bio-oil before the reactive species are
2019). Hydrotreatment upgrading is becoming a major refinery process polymerized. Although the concentrations of CO2 and CO are highly
for reducing the oxygen content and producing high-quality biofuels associated with hydrogen partial pressure (Stummann et al., 2019c),
(Mahesh et al., 2021). water–gas shift reaction also has effect on the concentrations of CO2 and
Hydrotreatment is conducted mostly under high hydrogen pressures CO, which can be catalyzed by the catalysts employed for FHP. There­
(100–170 bar), temperatures at 310–375 ◦ C, and low space velocities fore, it could be hard to experimentally prove the effects of decarbox­
(0.1–0.2 of LHSV, liquid hourly space velocity) in the presence of ylation and decarbonylation reactions on FHP.
catalyst. The use of catalysts can enhance hydrogenation of the Hydrogenation and HDO reactions also occur in FHP because
oxygenated compounds produced from the pyrolysis, leading to a higher hydrotreating catalysts are often employed for FHP (Stummann et al.,
H/C ratio of bio-oil (Ahmed et al., 2020). For example, while a typical 2018). They are likely to be the most important reactions in FHP because
effective H/C ratio of biomass ranges from 0 to 0.3, in situ hydrogena­ they substantially contribute to the decrease in oxygen content of the
tion over a Ru/C catalyst leads to a bio-oil having the ratio of > 1 (Huang bio-oil obtained (Kim et al., 2019a). In HDO, oxygen contained in
et al., 2016). This process, however, produces a large quantity of coke, organic species is reacted with hydrogen to form water. Thus, there is no
which deactivates the catalyst active sites and clogs the reactor, result­ carbon loss in removing the oxygen, and it allows a high yield of carbon
ing in additional cost to regenerate the catalyst (Wu et al., 2020). in bio-oil at the expense of a high hydrogen consumption compared with
Hydropyrolysis (HP) is a pyrolysis process performed under H2 at­ decarboxylation and decarbonylation. HDO of oxygenated aromatics (e.
mosphere, developed to obtain high quality liquid hydrocarbon fuels g., phenolic compounds) occurs via deoxygenation, retaining aroma­
while increasing the efficiency of hydrogen utilization. Fast hydro­ ticity, and hydrogenation; these reactions saturate the aromatic ring
pyrolysis (FHP) process to produce hydrocarbon products such as ole­ before the deoxygenation (Liu et al., 2021a). For HDO reactions, a high
fins, cycloalkanes, and alkanes, has recently been investigated (Resende, hydrogen pressure (e.g., >50 bar) is often needed to operate the process
2016). Even though recent research has suggested that FHP is a prom­ at a high temperature (e.g., >250 ◦ C) (Kim et al., 2019a). Major reaction
ising way to produce liquid hydrocarbon fuels from renewable resources pathways involved in HDO are highly dependent upon the components
such as biomass, only limited number of studies is available, and the HP of catalyst (Kim et al., 2019a). Jin et al. recently reported a catalytic
process is not well-understood. In this regard, this review focuses on the system comprising Mo-promoted Ir/SiO2 and HZSM-5 for the production
effectiveness of HP process for the production of hydrocarbon chemicals of C5 and C6 alkanes from cellulose via hydrogenation and hydro­
and provides a comparison of HP with conventional pyrolysis. The genolysis (Jin et al., 2020).
application of catalytic fast hydropyrolysis (CFHP) compared to con­ When using molybdenum sulfide-based catalysts for FHP of biomass,
ventional catalytic fast pyrolysis (CFP) is also discussed. The challenges H2S could be produced during the FHP process. In such cases, recom­
in hydrocarbon production via HP process are discussed, and the future bination between olefins and H2S takes place resulting in generation of
perspectives are also provided. organosulfur compounds such as thiols. Thiols can be further converted
into thiophenes via acid catalyzed reaction with carbenium ions
2. Mechanism of fast hydropyrolysis (Leflaive et al., 2002). Given that the incorporation of sulfur is catalyzed
by acid sites, the use of a catalyst having a low acid density might help
The initial reactions taking place in FHP should be similar to that reducing the S content of bio-oil produced from the FHP. Recombination
occurred in conventional fast pyrolysis, which are mostly associated reactions increase the S content in bio-oil, while the S content can be

2
S. Oh et al. Bioresource Technology 342 (2021) 126067

Fig. 1. Reaction chemistries and mechanisms for conventional fast pyrolysis, non-catalytic fast hydropyrolysis, and catalytic fast hydropyrolysis of biomass samples
(abbreviations inside the figure: HCR – hydrocracking; HDO – hydrodeoxygenation; DDO - direct deoxidation; DCO – decarboxylation; DME – demethylation; MT –
transmethylation; DE – deoxygenation extent; HLE – hydrogen-loss extent). Reprinted from (He et al., 2020), Copyright (2019), with permission from Elsevier.

readily reduced through subsequent hydrotreating of the bio-oil. 3. Non-catalytic fast hydropyrolysis
Occurrences of the aforementioned reactions are highly influenced
by temperature and pressure in FHP process, thereby affecting the Among the diverse biomass to liquid conversion processes, fast py­
product selectivities (Venkatesan et al., 2020). For example, an increase rolysis is one of the most preferred process. During fast pyrolysis process,
in temperature increases the ratio of aromatic/aliphatic hydrocarbons. the feedstock is converted to pyrolytic vapors via rapid heating for a
At low temperatures, hydrogenation of unsaturated double bonds (e.g., short residence time, followed by the pyrolytic vapors quenched to form
olefins, aldehydes, and ketones) occurs to stabilize bio-oil, whereas HDO liquid bio-oil (Liu et al., 2021b). The short residence time and rapid
takes place at higher temperatures to eliminate furanic and phenolic quenching could minimize the secondary reactions and generate the
oxygen-containing compounds (Wang et al., 2013). As seen in Fig. 2, maximum amount of liquid bio-oil by limiting the gas production
olefins, aldehydes, and ketones are hydrogenated to alcohols at low (Dabros et al., 2018). Bio-oil can be used widely for heating, generating
temperatures. Alcohols then undergo both catalytic hydrogenation and electricity, or producing value-added chemicals (Wang et al., 2021). On
thermal dehydration between 150 ◦ C and 200 ◦ C to form olefins. Car­ the other hand, some features of bio-oil, such as low energy density, high
boxylic groups, phenolic ethers, and phenols are reactive between oxygen content (ca. 40 wt%), high water content (ca. 20 wt%), high
300 ◦ C and 350 ◦ C. As hydrogen partial pressure increases, the ratio of acidity (acid number:100–200 mg of KOH per g of bio-oil), high vis­
aromatic/aliphatic hydrocarbons decreases. Hydrogen partial pressure cosity, instability, and immiscibility with petroleum, make the bio-oil
also plays a critical role in transforming oxygen-rich pyrolytic volatiles unsuitable for use as transportation fuel (Marker et al., 2012). Given
to hydrocarbons. An increase in the proportion of H2O and light non- that the disadvantages of bio-oil are generally due to the presence of
condensable compounds such as CO, CO2, and CH4 proves the occur­ oxygenated compounds (e.g., acids and aldehydes), bio-oil obtained
rence of deoxygenation via dehydration, decarboxylation, and decar­ from fast pyrolysis requires further upgrading via hydrotreatment (Kong
bonylation reactions. et al., 2020). According to a previous study, the H/C ratio of the most
deoxygenated bio-oil is approximately unity, and this value is similar to
that of coal and better than that of petroleum blends. Therefore, in-situ
or ex-situ hydrotreatment can improve the bio-oil composition (Gamliel

3
S. Oh et al. Bioresource Technology 342 (2021) 126067

and the enhancement in the CH4 yield was most substantial with a
sweeping-gas flow rate at 150 mL min− 1. The CH4 was produced mainly
through hydrogenation of tar. The presence of alkaline species promoted
the hydrogenation reaction. In Table 1, the advantages and disadvan­
tages of conventional non-catalytic fast pyrolysis and NCFHP are
compared.
FHP produces two phases of liquid products, referred to as the top
organic phase (target products consisting of a hydrocarbon mixture) and
the aqueous phase (based on the dehydrated water), as well as char and
gases (Marker et al., 2012; Resende, 2016). The organic phase obtained
from FHP has a relatively low oxygen level and is of better quality than
the bio-oil produced from conventional fast pyrolysis under an inert
atmosphere (Balagurumurthy & Bhaskar, 2014). NCFHP results in the
generation of large amounts of CH4, CO, and CO2 gases as by-products
with a tar-like char (Venkatakrishnan et al., 2014) because of insuffi­
cient hydrogenation of the oxygenated volatiles. The level of deoxy­
genation of the products, especially the targeted liquid phase, varies
with the operating conditions (e.g., reaction temperature, hydrogen
pressure, and the type of catalyst). For example, hydrogen can react with
pyrolytic volatiles under harsh conditions (above 500 ◦ C and 30 bar) to
generate low molecular weight hydrocarbons (i.e., the cracking reaction
is promoted by increasing hydrogen pressure) (Dayton et al., 2013).
Previous study reported that the HP of cellulose at high temperatures
(580 ◦ C) combined with high hydrogen pressures (50 bar) suppressed
Fig. 2. Reactivity scale of oxygenated groups in bio-oil according to tempera­ the formation of light oxygenated compounds (e.g., glycoaldehyde and
ture. Reprinted from (Wang et al., 2013), Copyright (2013) American Chemi­ formic acid) (Venkatakrishnan et al., 2014).
cal Society. Table 2 lists the representative experimental conditions and the re­
sults of NCFHP. It can be found that the yields of pyrolytic products vary
et al., 2018). depending upon NCFHP temperature and hydrogen partial pressure.
HP was initially studied for coal (Canel et al., 2005; Chareonpanich This indicates that the temperature and hydrogen partial pressure affect
et al., 2002), and then extended to produce hydrocarbons from biomass the extent of devolatilization and stabilization of volatiles during the
(Galiasso et al., 2014; Onay et al., 2006). HP can be conducted using NCFHP of biomass.
only hydrogen gas, or with hydrogen diluted in an inert gas (Ven­
katakrishnan et al., 2014). Under a hydrogen atmosphere, the free 4. Catalytic fast hydropyrolysis
radicals and olefins causing bio-oil instability are not formed due to the
capping of reactive volatile intermediates by hydrogen radicals (Melli­ During CFP, the components of biomass (e.g., cellulose, hemicellu­
gan et al., 2013; Thangalazhy-Gopakumar et al., 2012). The reaction of lose, and lignin) are degraded to form vapors at temperatures of
biomass with hydrogen during HP at elevated temperatures can produce 300–600 ◦ C under an inert atmosphere. CFP can be performed in either
a hydrocarbon-rich intermediate with a low oxygen content (Dayton two units (ex situ) or single unit (in situ) pyrolysis reactors (Chandler &
et al., 2013). The hydrogen pressure required in HP (around 30–50 bar) Resende, 2019). When CFP is carried out in a typical fluidized bed
is typically much lower than that required in bio-oil hydrotreatment reactor, the bed material inside the reactor can act as a heat-transfer
(over 100 bar is required). HP is generally performed at 30 bar of material and also catalyze the various reactions simultaneously (Dab­
hydrogen pressure. The partial hydrogen pressures are also recorded, ros et al., 2018). Typically, CFP is accomplished under atmospheric
and>50 bar of hydrogen pressure has also been studied (Ven­ conditions with zeolite catalysts (HZSM-5 (Li et al., 2020; Xu et al.,
katakrishnan et al., 2014), and bio-oil hydrotreatment requires high
hydrogen pressures (100–170 bar) (Marker et al., 2012; Marker et al.,
Table 1
2014).
Comparison of advantages and disadvantages of different pyrolysis types.
Apart from typical hydrotreatment (conducted with high pressure at
around 100–170 bar to reduce coke formation and plugging problem Pyrolysis type Advantage Disadvantage

during re-vaporization of bio-oil), fast pyrolysis under hydrogen atmo­ Non-catalytic - Relatively simple reactor design - High oxygen content in
sphere (i.e., FHP) was developed to overcome the problems mentioned fast pyrolysis bio-oil
- Low selectivity toward
above and produce hydrocarbon-based liquid products. FHP, whereby
target products
the organic material is decomposed rapidly under a hydrogen atmo­ CFP - High selectivity toward target - Catalyst deactivation
sphere (approximately 500 ◦ C s− 1), has been studied as the hydro­ products (e.g., BTX)
pyrolysis process uses a long residence time and low heating rates - Lower downstream hydrotreating
cost than non-catalytic fast
(Venkatakrishnan et al., 2015). Similar with fast pyrolysis, FHP can be
pyrolysis
performed with fluidized bed reactors (Dayton et al., 2013; Marker et al., NCFHP - Biofuels with low oxygen content - Need for expensive
2012; Marker et al., 2014), cyclone reactors (Venkatakrishnan et al., and high quality pyrolysis agent such as
2014), and micropyrolysis systems such as Py-GC/MS (pyroprobe-gas - Eliminating the need for H2
chromatography/mass spectrometry) (Jan et al., 2015; Thangalazhy- reheating
- Lessening undesired reactions
Gopakumar et al., 2012), which allows short residence times (<2 s).
- Producing biofuel with low
Tian et al. has recently suggested a partial NCFHP process operating oxygen content and high quality
at temperatures between 700 ◦ C and 900 ◦ C under 0% to 30% hydrogen CFHP - Higher selectivity toward - Need for expensive
atmosphere (Tian et al., 2021). An increase in the H2 concentration in hydrocarbons than NCFHP pyrolysis agent such as
pyrolysis medium led to enhancing the yields of CH4, C2H6, and light tar, - Higher deoxygenation extent then H2
NCFHP - Catalyst deactivation

4
S. Oh et al. Bioresource Technology 342 (2021) 126067

Table 2 decreased coke formation due to the combination of Ni metal sites and
Summary of non-catalytic fast hydropyrolysis studies reported in earlier zeolite acid sites (Iliopoulou et al., 2012; Valle et al., 2010; Veses et al.,
literature. 2015). As a transition metal, Ni can promote oligomerization and
Feedstock Conditions Performance Ref. dehydration (Iliopoulou et al., 2012), as well as increasing the hydro­
Pinewood and 300–700 ◦ C; High pressure (40–70 bar) (Galiasso et al.,
thermal stability and cost-effectiveness of the catalyst (Gamliel et al.,
lignin 70 bar cracked much of the 2014) 2017a). The HZSM-5 catalyst is also known as effective in producing
lighter compounds compatible hydrocarbons, such as aromatics (e.g., benzene, toluene, and
Eucalyptus 450 ◦ C; 100 s Deeper bed (high pressure) (Onay et al., xylene), while these aromatics may form coke during reforming at low
wood waste produced a lighter, less 2006)
temperature (<200 ◦ C) or remain inside the zeolite pores. Those coke
oxygenated, and more
stable tars/oils formations should be prohibited since the coke could block the active
Sugarcane 600 ◦ C; 70 Almost 30 wt% liquid (Venkatakrishnan sites of the catalyst and cause its deactivation (Carlson et al., 2011). CFP
bagasse bar product is obtained et al., 2014) of sawdust at the temperature ranging from 500 to 670 ◦ C with a weight
Sunflower 400–700 ◦ C High hydrogen pressure (Melligan et al., hourly space velocity (WHSV) of 0.1–1.7 h− 1 resulted in decreased
pressed (>50 bar) increased the oil 2013)
bagasse yields (~10 wt%)
toluene and xylene production, whereas naphthalene production
Mixed wood, 350–480 ◦ C; 20.6–46.3 wt% liquid (Mohan et al., increased with increasing WHSV. A maximum aromatic hydrocarbon
bagasse, 14–35 bar yield with over 90% 2006) yield of 14% was obtained at 600 ◦ C and 0.1 h− 1 (Carlson et al., 2011).
corn stover, carbon recovery The formation of a stable liquid product under inert conditions (i.e.,
microalgae,
pyrolysis without using hydrogen as carrier gas) is an advantage of CFP,
seaweed
Euphorbia 400–650 ◦ C; Liquid product yield of (Marker et al., but coke formation is a major disadvantage of CFP. For example, only
rigida, 50 bar 30–40 wt%; maximum 2014) 10–30 wt% of carbon could be recovered as a liquid product, with
sunflower yield achieved at 520 ◦ C approximately 30 wt% of the carbon loss as coke deposited on the
oil, catalyst active sites (Agblevor et al., 2010; Resende, 2016). The oxygen
industrial
content remained at a significant level between 10 wt% and 20 wt%
waste, pure
cellulose after the CFP process (Stummann et al., 2020b; Venderbosch, 2015).
Pine wood 550 ◦ C; 5.5 Biomass to catalyst ratio (Jan et al., 2015) Therefore, to overcome this limitation, introduction of hydrogen gas
chips bar from 1:4 to 1:9 reduced the into the pyrolysis process was suggested (Gamliel et al., 2018).
oxygenated compounds
FHP in the presence of a HDO catalyst for deoxygenation and hy­
and increased the aromatic
compounds drogenation under a partial hydrogen atmosphere is termed CFHP (Chen
Loblolly pine 375–450 ◦ C; Increased hydrogen partial (Anuar Sharuddin et al., 2020b; Linck et al., 2014). The main effect of using hydrogen in
3.4–20 bar pressure promoted liquid et al., 2016) FHP is to reduce the coke yield and increase the generation of liquid
production hydrocarbons and gaseous olefins. The addition of a catalyst favored the
Pine sawdust, 700–900 ◦ C; CH4, C2H6, and light tar (Tian et al., 2021)
deoxygenation pathway compared to the non-catalytic system, resulting
reeds, 0–30% H2 yields increased with
Sargassum increasing H2 in relatively low oxygen content in bio-oil (4 wt%) (Stummann et al.,
horneri concentration 2019b), and the generation of aromatic hydrocarbons as the primary
Pine sawdust 400–600 ◦ C; Higher yield of (Venkatesan product in the organic liquid phase. Table 1 compares the advantages
1–20 bar condensables than typical et al., 2020)
and disadvantages of CFP and CFHP are compared.
pyrolysis; More
hydrocarbons at high H2
CFHP may unavoidably lead to coke production (even small amounts
partial pressures; more of coke was produced at the optimal condition), which deactivates the
pyrolysates and less char catalyst activity (Marker et al., 2012; Resende, 2016). For example,
were produced as CFHP with HZSM-5 at 400 ◦ C and 35 bar could produce 8.9 gallon of
temperature increased
liquid hydrocarbons (mostly consisted of BTX) per ton of biomass,
Poplar 500 ◦ C; It had a significant (He et al., 2020)
sawdust, atmospheric hydrodeoxygenation whereas 6.0 gallon of liquid hydrocarbons per ton of biomass was ob­
rice husk pressure effect, compared with tained with CFP (Chandler & Resende, 2019).
conventional fast pyrolysis For the CFHP of lignin model compounds, ReOx/MCM-41, ZrO2, and
ReOx/ZrO2 supports could effectively reduce coke formation compared
with HBeta (Sirous-Rezaei et al., 2018; Sirous-Rezaei & Park, 2020).
2021; Zhang et al., 2019)) or alumina-based catalysts (Foong et al.,
During the CFHP process, oxygen is removed by conversion to water or
2021; Mante et al., 2016; Nguyen et al., 2017). When catalytic fast py­
gases (e.g., CO and CO2), which minimizes the undesirable reactions,
rolysis was performed with the zeolite, approximately 25% carbon re­
such as acid catalyzed polymerization, aromatization, and coking re­
covery in the form of hydrocarbons, mainly as olefins and aromatics,
actions, promoting the hydrocarbon production by addition of hydrogen
was obtained (Lappas et al., 2012). In particular, biomass was pyrolyzed
to the hydrocarbon structure using the active catalyst (Balagurumurthy
to produce hydrocarbon ca. 15–25 wt% using ZSM-5 as a catalyst (Liu
et al., 2013). Furthermore, the active catalyst used in CFHP can promote
et al., 2014). To enhance the hydrocarbon yield and reduce the oxygen
the addition of hydrogen to the hydrocarbon structure, resulting in hy­
content, ZSM-5 was modified and utilized in the CFP (Che et al., 2019;
drocarbon production (Balagurumurthy et al., 2013). During those un­
Dai et al., 2019; Shang et al., 2021). The pyrolytic vapors were hydro­
desirable reactions, the intermediates may undergo polymerization and
treated immediately in the catalyst bed reactor (placed after the fast
generate carbon-based coke. On the other hand, the hydrogen, added to
pyrolysis reactor) to achieve a certain level of deoxygenation (Jiang
the CFHP process, can cap the active sites of the intermediates and
et al., 2020; Pirbazari et al., 2019; Wang et al., 2019). Therefore, the
reduce coke formation (Melligan et al., 2013; Thangalazhy-Gopakumar
selection or design of an appropriate catalyst for the CFP process is
et al., 2012). To enhance the deoxygenation reaction, elevated pressure
critical in catalytic fast pyrolysis to produce high yield of hydrocarbons.
was applied to the CFHP process. This elevated pressure resulted in
The catalyst used in CFP must have a capacity to remove the oxygen-
deoxygenation reaction, with the high energy recovery and a highly
containing functional groups and add hydrogen atoms into the feedstock
deoxygenated liquid product (Venderbosch, 2015).
(Alonso et al., 2012; Huber et al., 2006; Kloekhorst et al., 2015). The CFP
According to a model compound study, anisole was converted to
process can produce more aromatic hydrocarbon than non-catalytic
cyclohexane over Ni/ZSM-5 at low temperatures and high pressures,
process (Dabros et al., 2018; Venderbosch, 2015). Studies on CFP with
whereas high temperatures could produce naphthalenes or poly­
Ni/ZSM-5 reported an enhanced yield of aromatic compounds and
aromatic hydrocarbons via polymerization (Gamliel et al., 2018). When

5
S. Oh et al. Bioresource Technology 342 (2021) 126067

pine was subjected to catalytic fast hydropyrolysis at 650 ◦ C and 28 bar 5.5 bar) produced a similar yield of aromatic hydrocarbons under a
with pyroprobe GC/MS using various metals supported on ZSM-5, it was helium atmosphere, even at higher pressures of about 27.5 bar (Resende,
suggested that the metals catalyzed the hydrogenation process, whereas 2016; Thangalazhy-Gopakumar et al., 2012). The selectivity toward
ZSM-5, which has a number of acid sites, notably influenced the deox­ toluene increased with enhancing the pressure due to the occurrence of
ygenation reaction (Santana Junior et al., 2019; Thangalazhy- simultaneous cracking and hydrogenation, whereas the production of
Gopakumar et al., 2012). When the CFHP experiments were conduct­ m/p-xylenes decreased. This result suggests that the reaction between
ed with laboratory scale pyroprobe GC/MS, the yield of aromatic hy­ the methyl group of m/p-xylenes and hydrogen may form CH4 gas
drocarbons increased at 600 ◦ C and 31 bar with M/ZSM-5 (M = Ni, Pd, (Resende, 2016; Thangalazhy-Gopakumar et al., 2012).
or Ru) (Gamliel et al., 2017a; Gamliel et al., 2017b; Xue et al., 2020), The CFHP of loblolly pine using a pre-reduced commercial hydro­
and at 650 ◦ C and 17 bar with Pd/HZSM-5 (Jan et al., 2015), compared treatment catalyst placed in a fluidized bed reactor at 375 ◦ C and 3 bar
to the case of HZSM-5 (Dabros et al., 2018). Previous research using two- of hydrogen pressure reached a 90.3% degree of deoxygenation, despite
stage upgrading of CFHP followed by a hydrotreatment over Pd/ZSM-5 a liquid product yield of only 10 wt% (Dayton et al., 2013). The hy­
concluded that single catalytic fast hydropyrolysis is thermodynamically drocarbon distribution of the liquid product depends significantly on the
insufficient for alkane formation and highlighted the need for additional bed materials used in the pyrolysis reactor. When MgAl2O4 or a mixture
ex-situ upgrading process (Gamliel et al., 2018; Gamliel et al., 2017b; of HZSM-5 and Al2O3 was used as the bed material, the CFHP of loblolly
Jan et al., 2015; Perkins et al., 2019). The CFHP of silvergrass at 600 ◦ C, pine produced char (18.7–21.1 wt% yield), CO (18.9 wt% yield), and
followed by fixed-bed upgrading at low temperatures over Ni/ZSM-5 CO2 (20.0 wt% yield), as well as 17.8–20.4 wt% of condensed organics
could effectively produce alkanes with increasing hydrogen pressure over C4 as a liquid product (Dayton et al., 2013). In addition, CoMo/
(Melligan et al., 2012; Melligan et al., 2013). M/H-ZSM-5 (M = 5 wt% MgAl2O4 and a mixture of NiMo/zeolite and Al2O3 could decrease the
Ni, Co, or Mo or 0.5 wt% Pt) converted lignin effectively to aromatic char yield significantly to 11.4–13.1 wt%; the condensed organics
hydrocarbons at hydrogen pressures of>20 bar (Thangalazhy-Gopaku­ increased to 21.5 wt% using CoMo/MgAl2O4 and to 24.0 wt% using a
mar et al., 2012). At 1 wt% of Pd loading, Pd/H-ZSM-5 produced 44% NiMo/zeolite and Al2O3 mixture (Stummann et al., 2020b). Sulfided
more aromatic hydrocarbons than those obtained with H-ZSM-5 at CoMo and NiMo could convert beech wood into upgraded biofuel
650 ◦ C and 17.2 bar, with a high catalyst to lignin ratio as 20:1 (Jan (18.7–21.5 wt%) with low oxygen level (<0.01 wt%) (Stummann et al.,
et al., 2015; Resende, 2016). On the other hand, CFHP using HZSM-5 2019c). Table 3 provides an overview of various CFHP processes
under mild conditions (catalyst to biomass ratio of 5:1 and pressure of available in earlier literature.

Table 3
Summary of previous studies on catalytic fast hydropyrolysis.
Feedstock Conditions Reactor Catalyst Ref.

Rice husk 450–650 ◦ C; atmospheric Fluidized bed Ni/Al2O3, Ni/LY, dolomite, CoMo/Al2O3 (Meesuk et al., 2011;
pressure Resende, 2016)
Rice husk 700 ◦ C Py-GC/MS Thermoplastic bituminous coal (Zhang et al., 2018)
Pine wood 650 ◦ C; 5.5–28 bar Py-GC/MS HZSM-5, Co/ZSM-5, Ni/ZSM-5, Pt/ZSM-5 (Thangalazhy-
Gopakumar et al., 2012)
Beech wood 450 ◦ C; 26 bar Fluidized bed Mo/MgAl2O4, CoMo/MgAl2O4, NiMo/ (Stummann et al., 2019d)
MgAl2O4
Poplar sawdust 450–550 C; 3–5 s

Py-GC/MS Rh/ZrO2 (He et al., 2020)
Lignin 650 ◦ C and < 25.8 bar Py-GC/MS Pd/HZSM-5 (Jan et al., 2015)
m-cresol and Guaiacol 350 ◦ C; 1 bar Py-GC/MS NiReOx/CeO2, NiReOx/ZrO2, NiReOx/ZrCeO2 (Sirous-Rezaei et al.,
2019)
Woody residue 375–450 ◦ C; 3.4–20 bar Fluidized bed Commercial hydrotreating catalyst, (Dayton et al., 2013)
commercial FCC catalyst
Green algae (Oscillatoria) 550 ◦ C; 1 bar Tubular TiO2 and ZnO (Kawale & Kishore, 2019)
Algae 260–450 ◦ C Fluidized bed MoS2 (Lee et al., 2020c; Love
et al., 2005)
Algal consortium (Chlorella 150–300 ◦ C; 25 bar Batch K2CO3 (Choudhary et al., 2020)
and Phormidium sp.)
Seaweed (Enteromorpha 550 ◦ C; atmospheric pressure Fixed bed MCM-41, ZSM-5 (Hu et al., 2021)
clathrate)
Lignin extracted from corn 600 ◦ C; atmospheric pressure Two-stage consisting of non- HZSM-5, MoO3/ZSM-5, Ni/ZSM-5 (Xue et al., 2020)
stover, loblolly pine, red catalytic reactor and fixed-bed
oak reactor in series
Pine sawdust 400–600 ◦ C; 1–20 bar Packed bed Sulfided metal catalysts (Venkatesan et al., 2020)
Beech wood ~450 ◦ C; >90 vol% H2 in N2 Fluid bed CoMo/MgAl2O4, MgAl2O4, olivine sand, bog (Stummann et al., 2020b)
iron, a HZSM-5/Al2O3 mixture, NiMO/HZSM-
5/Al2O3 mixture
Beech wood 450 ◦ C; 94 vol% H2 in N2 Fluid bed CoMo/MgAl2O4, CoMo/HZSM-5/Al2O3 (Stummann et al., 2020a)
Kraft lignin First zone: 600 ◦ C, second zone: Tandem micro-reactor with two HY, HBeta, Fe/HBeta, FeReOx/MCM-41, Fe/ (Sirous-Rezaei & Park,
350–550 ◦ C; atmospheric zones ZrO2, ReOx/ZrO2, FeReOx/ZrO2, NiReOx/ 2020)
pressure ZrO2, PdReOx/ZrO2
Poplar sawdust, rice husk 500 ◦ C; atmospheric pressure Py-GC/MS Rh/ZrO2 (He et al., 2020)
Mixture of pine sawdust/ 360–420 ◦ C; 5–20 bar Packed bed HZSM-5 (Chen et al., 2020b)
HDPE
Arundo donax 400 ◦ C; 35 bar Fluidized bed NiMo/HZSM-5 (Chandler & Resende,
2019; Chandler et al.,
2020)
Beech wood ~450 ◦ C; >90 vol% H2 in N2 Fluid bed Sulfided NiMo/MgAl2O4 (Stummann et al., 2019c)
Kraft lignin 450–650 ◦ C; 6.9 bar Py-GC/MS ZSM-5, HY-340 (Santana Junior et al.,
2019)
Microalgae > 120 ◦ C Batch Mo2C (Azizi et al., 2018; Chang
et al., 2015)

6
S. Oh et al. Bioresource Technology 342 (2021) 126067

A catalytic route to produce liquid alkane from biomass (giant cane) bed. On the other hand, the operating conditions of HP and HDO may be
using a NiMo/H-ZSM-5 catalyst was suggested, involving methane for­ dissimilar; the reaction temperature is higher in HP (480–580 ◦ C) than
mation promoted by nickel, oligomerization of methane to aromatic HDO (300–375 ◦ C), but the pressure is lower in HP (<70 bar) than HDO
compounds catalyzed by molybdenum, and hydrogenation of aromatic (≥70 bar) (Nguyen & Clausen, 2019).
compounds to cycloalkanes catalyzed by nickel (Chandler et al., 2020). Compared to CFHP, NCFHP can produce a high yield of hydrocar­
The highest hydrocarbon yield was 7.9 wt% at 400 ◦ C under 35 bar H2, bons without pronounced catalyst deactivation (Resende, 2016).
which corresponds to 15 gallons of hydrocarbons per ton of biomass. Therefore, H2Bioil, a process combining NCFHP, vapor-phase catalytic
The increase in Ni loading in the catalyst suppressed coke formation at hydrotreatment, and HDO, has been proposed as an alternative process
the expense of an increase in the yield of C2–C3 alkanes. for producing hydrocarbon fuels (Agrawal & Singh, 2009; Singh et al.,
To produce high-quality bio-oil, reduce the energy load on coal, and 2010). In this process, biomass is hydropyrolyzed rapidly at around
utilize waste plastic, the co-pyrolysis of biomass and coal (or waste 480–580 ◦ C and the vapor products are then hydrodeoxygenated at
plastics) has been performed and recommended (Klose & Stuke, 1993; 300–375 ◦ C under high hydrogen pressure (~200 bar), resulting in high
Sharma & Ghoshal, 2010). On the other hand, the co-pyrolysis of efficiency hydrocarbon production (Nguyen & Clausen, 2019; Ven­
biomass and coal could form heavy tar that can lead to operational katakrishnan et al., 2015). The reactive oxygenated compounds could be
problems, such as pipe blocking and tar aerosol formation. The co-HP of upgraded effectively via HDO before undesirable secondary reactions
coal/synthetic polymer mixtures (Sharypov et al., 2007; Yasuda et al., could occur, and this process could result in the strikingly reduced ox­
2014) and biomass/synthetic polymer mixtures (Pinto et al., 2016; ygen containing compounds (Venkatakrishnan et al., 2015). This H2Bioil
Sharypov et al., 2006) is expected to improve the hydrocarbon yield. process is estimated to show high carbon recovery (up to 70%) as well as
The co-HP of rice husk and thermoplastic bituminous coal improved the a high energy efficiency up to 75%. When the H2Bioil process was
higher heating value and enhanced the BTX (benzene, toluene, and performed with cellulose or with biomass feedstock (e.g., poplar), the
xylene) yield, whereas the production of corrosive acids, such as carbon recoveries for C1–C8+ hydrocarbons reached up to 73% with
methanoic acid, acetic acid, and propanoic acid, decreased by 89% cellulose and 54% with poplar while the yield of C4+ hydrocarbons
under 50 bar of hydrogen (Zhang et al., 2018). within the liquid fuel range was up to 55% and 32%, respectively
He et al. used a modified lab-scale micro-pyrolyzer system to (Venkatakrishnan et al., 2015). This was notably a higher amount of
compare conventional fast pyrolysis, NCFHP, and CFHP over a Rh/ZrO2 hydrocarbons (17.8–20.4 wt%) compared with that produced from
catalyst (Rh loading of 0.5 wt%) (He et al., 2020). In terms of the CFHP (Dayton et al., 2013). According to the economic analysis, this
selectivity towards hydrocarbons and the extent of deoxygenation, H2Bioil is also economically attractive in both small and large scale
CFHP showed the best performance among the three processes. For process (USD 0.62 to 0.73 per liter), since it can result in a high yield of
poplar sawdust, CFHP showed a 49.1% selectivity towards hydrocar­ hydrocarbons that can be employed as transportation fuel without
bons, while NCFHP and conventional fast pyrolysis showed ~ 10% and further treatment after HP (Zhang et al., 2017).
~ 1% selectivities, respectively. The extent of deoxygenation decreased Another integrated process developed by the Gas Technology Insti­
in an order of CFHP > NCFHP > conventional fast pyrolysis. When using tute (GTI), IH2®, consists of bubbling fluidized bed reactor HP at
the feedstock with a high ash content (e.g., rice husk), NCFHP showed 335–480 ◦ C, followed by downstream HDO (345–400 ◦ C) of the vapor
the performance comparable to CFHP, most likely because the mineral type products in a fixed bed reactor, with a 20–30 bar of hydrogen
elements contained in ash bring about unexpected and uncontrollable pressure in both reactors to sustain deoxygenation reactions (Marker
catalytic effects (Carpenter et al., 2014; Li et al., 2016). et al., 2012; Marker et al., 2014). This IH2® pilot unit could continu­
Based on the results reported in various studies, the employment of a ously convert 50 kg per day of feedstock, with long term performance of
right catalyst for FHP process is essential to achieve a high carbon re­ catalyst (endure almost 700 h of operation) in 2012, and the process
covery from raw biomass with a low oxygen content. However, catalyst expanded to 2,000 metric tons per day of wood. Through this process,
deactivation is a major challenge for the CFHP, which is possibly caused the oxygen content of the liquid product was typically<0.5 wt% and the
by transfer of alkali metals from biomass feedstock to catalyst (Stum­ energy recovery was approximately 65% with a liquid product yield of
mann et al., 2019a), catalyst attrition/entrainment (Dayton et al., 2016), 26–30 wt% (Dabros et al., 2018). Non-condensable hydrocarbons such a
and metal poisoning. In addition, crushed catalyst due to attrition is (e.g., CH4, C2H6, and C3H8) produced from the IH2® process are not
impossible to separate from char after the HP. This might constitute a suitable for transportation fuel. Therefore, those hydrocarbons were sent
serious challenge because some HP catalysts (e.g., sulfide catalysts) are to a steam reformer as a vapor phase to produce H2 and CO2, which
toxic, thereby the char unseparated from the used catalyst being toxic could then be sent back into the process (Dabros et al., 2018). The liquid
waste. Therefore, the development of stable and non-toxic catalysts for products obtained via the IH2® process can have higher contents of
FHP should be more encouraged. carbon (C: 85–88 wt%) and hydrogen (H: 12–14 wt%) and lower oxygen
content (O: <2.2 wt%) than those obtained via HDO (C: 76–87 wt%; H:
5. HFP combined with hydrotreatment process 9.7–13 wt%; O: 0.02–14.2 wt%) or CFP (C: 77–90 wt%; H: 9.6–12 wt%;
O: 0.5–14 wt%) (Dabros et al., 2018; Marker et al., 2012; Marker et al.,
Since C6 to C8 aromatic hydrocarbons can be used as high-octane 2014; Singh et al., 2010). The higher heating value (HHV) of the liquid
gasoline additives, transportation fuel itself, and intermediates of product produced via the IH2® process (55 MJ kg− 1) is also higher than
manufacturing polymers. HDO is an effective upgrading process because that of the liquid product produced via HDO process (41–46 MJ kg− 1)
it removes oxygen and leads to an atomic H/C ratio (1.75) of bio-oil that (Meesuk et al., 2011). From techno-economic assessment of the IH2®
is close to that of gasoline (Gamliel et al., 2018; Morvay & Gvozdenac, process, it was estimated that the product prices ranged from USD 0.43
2008). Direct HDO of biomass with multifunctional Pt/NbOPO4 catalyst to 0.50 per liter (Tan et al., 2014), whereas the techno-economic
yielded 28.1 wt% of liquid alkanes (Xia et al., 2016). Typically, the HDO assessment estimated the value of bio-oil produced from conventional
process was performed with bio-oil produced from fast pyrolysis, and fast pyrolysis-based process to be around USD 0.82 per liter (Wright
complete HDO required large amounts of hydrogen with high pressure. et al., 2010).
CFHP primarily produces aromatic hydrocarbons, but appropriate Xie et al. reported an integrated HP/hydrotreatment sequenced
conditions with hydrotreatment can also produce alkanes and naph­ batch process for the conversion of microalgae (Auxenochlorella pyr­
thalenes. FHP and CFHP can be combined with flowthrough vapor- enoidosa) and used engine oil into sulfur-free liquid fuel (Xie et al.,
phase HDO for effective upgrading in a fixed bed reactor (Dabros 2019). For the first step of the integrated process, a blend of microalgae
et al., 2018; Saraeian et al., 2019). The hydropyrolytic vapors from and used engine oil was processed at 400 ◦ C under 60 bar H2 for 4 h in
biomass are upgraded to hydrocarbons by HDO in a catalytic packed the presence of a Pt/C catalyst, resulting in production of light and

7
S. Oh et al. Bioresource Technology 342 (2021) 126067

heavy oil. The light oil was recovered and processed in the second step of fuels (energy use, CO2 and SO2 emissions, etc.) or encouraged in­
the integrated process at the same condition to that used in the first step. vestments related to the renewable energy (Abdmouleh et al., 2015). For
The light oil obtained in the second step was also recovered and pro­ instance, Australia (Renewable Energy Demonstration Program; REDP),
cessed repeatedly until the fourth step. The yields of light oil achieved China (Global Environmental Facility; GEF) and China Renewable En­
from four runs were between 83 wt% and 89 wt%. The contents of O, N, ergy Scale-Up Program; CRES), Indonesia (Integrated Micro hydro
and S decreased when the integrated process was applied. The light oil Development and Application Program; IMIDAP), Sri Lanka (Renewable
obtained after the fourth processing had fuel properties that are similar Energy for Rural Economic Development Project; RERED), Nepal
to gasoline (zero sulfur content, 2 ppm of nitrogen, acid number of 1.89, (Biogas Support Program; BSP), and Europe (Austria, Ireland, Portugal,
density of 0.8 kg m− 3, kinematic viscosity of 1.89 × 10− 3 m2 s− 1, and and Spain) provide financial support for renewable energy development
HHV of 46.4 MJ kg− 1). This study showed that HP process combined (Abdmouleh et al., 2015). According to IEA/IRENA (International En­
with hydrotreatment can be effective at producing gasoline-like fuels ergy Agency/International Renewable Energy Agency) Joint Policies
from biomass resources. and Measures database, nine countries (Angola, Bulgaria, El Salvador,
Denmark, Paraguay, Poland, Fiji, India, and France) implemented
6. Research needs and perspectives financial support policies toward bioenergy in force from 2010 to 2017.
Therefore, although few studies have already been available (Zupko,
FHP is a promising process to make renewable hydrocarbons that 2019), more studies into technoeconomic analysis, life cycle assess­
have a potential to replace fossil hydrocarbons. This process shows ments, as well as carbon and energy analysis for integrated HP and
beneficial features to overcome the disadvantages of conventional fast hydrotreatment processes should be conducted to make them more
pyrolysis (production of low-quality bio-oil) as well as catalytic hydro­ industrially viable.
treatment (requirement of high hydrogen pressure). The application of
CFHP shows potential to prolong the catalyst lifetime via suppression of 7. Conclusions
the coke formation as well as demonstrating high hydrocarbon selec­
tivity by minimizing the occurrence and extent of undesirable reactions. Herein, an effective bio-oil upgrading process, HP, is reviewed. It is a
Fast pyrolysis represents a thermochemical conversion process that process of pyrolysis operating under H2 environment. Although slow HP
can produce bio-oil and hydrocarbons (e.g., aromatic hydrocarbons such has been investigated for many years, FHP has been primarily reported
as benzene, toluene, ethylbenzene, and xylenes). Although conventional only over the last few years. Aromatic hydrocarbons are primary prod­
fast pyrolysis is commercially available, the utilization of bio-oil as a ucts of CFHP, while alkanes and cycloalkanes are also produced when an
boiler fuel or in industrial applications is restricted (Butler et al., 2011; ex-situ hydrotreating unit is applied. CFHP results in higher yields of
Huber et al., 2006; Perkins et al., 2018). As a solution to tackle the issue, hydrocarbons and slower catalyst deactivation than conventional CFP.
it was suggested that bio-oil is hydrotreated followed by co-processing This shows that FHP is a promising option to upgrade renewable alter­
with petroleum-derived hydrocarbons such as fuel oil (Lappas et al., native biofuels to more realistic fuels.
2009; Pinho et al., 2015); however, bio-oil is immiscible with such liquid
hydrocarbon due to its high oxygen content and causes coke formation
during its hydrotreating process (Talmadge et al., 2014). To optimize the Declaration of Competing Interest
production of high concentration of C6 to C8 aromatic hydrocarbon for
use as transportation fuel or gasoline additives, integrated processes The authors declare that they have no known competing financial
coupled with FHP and HDO (e.g., H2Bioil and IH2® processes) have been interests or personal relationships that could have appeared to influence
commercially developed. Compared with the technoeconomic analysis the work reported in this paper.
of bio-oil production, a commercial IH2® process is estimated to be
competitive. Acknowledgments
CFHP system could effectively enhance the product quality and
produce large amounts of hydrocarbons (e.g., aromatic hydrocarbons) This study was supported by the National Research Foundation of
which can be employed as a boiler fuel and a transportation fuel Korea (NRF) grant funded by the Korea government (MSIT) (No. NRF-
(Khodier et al., 2009; No, 2014). The addition of a catalytic conversion 2020M1A2A2079801, NRF-2021R1A2C3011274).
unit (e.g., CFHP and HDO) is required to obtain a high yield of bio-oil
with a low oxygen content. The formulation of the catalyst used in References
CFHP is essential, and issues related to catalyst deactivation (e.g.,
catalyst poisoning, coke formation, and leaching), catalyst regeneration, Abdmouleh, Z., Alammari, R.A.M., Gastli, A., 2015. Review of policies encouraging
renewable energy integration & best practices. Renew. Sust. Energ. Rev. 45,
as well as reactor breakdown, need to be addressed (Dabros et al., 2018). 249–262.
To maximize the hydrocarbon yield for utilization as a fuel or chemicals, Agblevor, F.A., Mante, O., Abdoulmoumine, N., McClung, R., 2010. Production of stable
it is important to design active, selective, and stable catalysts (Kim et al., biomass pyrolysis oils using fractional catalytic pyrolysis. Energy Fuel. 24 (7),
4087–4089.
2019b). The complex reaction mechanism of biomass HP or catalytic Agrawal, R., Singh, N.R., 2009. Synergistic routes to liquid fuel for a petroleum-deprived
conversion causes the process and product to differ in each case. future. AIChE J. 55 (7), 1898–1905.
Therefore, a detailed study, including the kinetic parameters depending Ahmad, M.S., Liu, C.-G., Nawaz, M., Tawab, A., Shen, X., Shen, B., Mehmood, M.A.,
2021. Elucidating the pyrolysis reaction mechanism of Calotropis procera and
on the different kinds of target products should be performed (Bala­ analysis of pyrolysis products to evaluate its potential for bioenergy and chemicals.
gurumurthy et al., 2013). Various catalysts apart from zeolite and HDO Bioresour. Technol. 322, 124545.
catalysts must be screened, and their corresponding reaction mecha­ Ahmed, M.H.M., Batalha, N., Mahmudul, H.M.D., Perkins, G., Konarova, M., 2020.
A review on advanced catalytic co-pyrolysis of biomass and hydrogen-rich feedstock:
nisms in each process need to be evaluated. Insights into synergistic effect, catalyst development and reaction mechanism.
The price of gasoline and diesel produced from conventional fast Bioresour. Technol. 310, 123457.
pyrolysis followed by hydrotreatment or hydrocracking was estimated Alonso, D.M., Wettstein, S.G., Dumesic, J.A., 2012. Bimetallic catalysts for upgrading of
biomass to fuels and chemicals. Chem. Soc. Rev. 41 (24), 8075–8098.
to be USD 0.54 per liter in 2007 based on the data from the Pacific
Anuar Sharuddin, S.D., Abnisa, F., Wan Daud, W.M.A., Aroua, M.K., 2016. A review on
Northwest National Laboratory (Jones et al., 2009). However, it is pyrolysis of plastic wastes. Energy Convers. Manag. 115, 308–326.
difficult to perform a direct comparison of renewable hydrocarbon fuels Azizi, K., Keshavarz Moraveji, M., Abedini Najafabadi, H., 2018. A review on bio-fuel
made via FHP with fossil hydrocarbon fuels because of the varied fuel production from microalgal biomass by using pyrolysis method. Renew. Sust. Energ.
Rev. 82, 3046–3059.
prices and different investments for each process (Nguyen & Clausen, Balagurumurthy, B., Bhaskar, T., 2014. Hydropyrolysis of lignocellulosic biomass: state
2019). Moreover, the existing policies have imposed a duty on fossil of the art review. Biomass Conv. Bioref. 4 (1), 67–75.

8
S. Oh et al. Bioresource Technology 342 (2021) 126067

Balagurumurthy, B., Oza, T.S., Bhaskar, T., Adhikari, D.K., 2013. Renewable Jiang, L., Wang, Y., Dai, L., Yu, Z., Wu, Q., Zhao, Y., Liu, Y., Ruan, R., Ke, L., Peng, Y.,
hydrocarbons through biomass hydropyrolysis process: challenges and Xia, D., Jiang, L., 2020. Integrating pyrolysis and ex-situ catalytic reforming by
opportunities. J. Mater. Cycles Waste Manag. 15 (1), 9–15. microwave heating to produce hydrocarbon-rich bio-oil from soybean soapstock.
Butler, E., Devlin, G., Meier, D., McDonnell, K., 2011. A review of recent laboratory Bioresour. Technol. 302, 122843.
research and commercial developments in fast pyrolysis and upgrading. Renew. Sust. Jin, L., Li, W., Liu, Q., Ma, L., Hu, C., Ogunbiyi, A.T., Wu, M., Zhang, Q.i., 2020. High
Energ. Rev. 15 (8), 4171–4186. performance of Mo-promoted Ir/SiO2 catalysts combined with HZSM-5 toward the
Canel, M., Mısırlıogˇlu, Z., Sınagˇ, A., 2005. Hydropyrolysis of a Turkish lignite conversion of cellulose to C5/C6 alkanes. Bioresour. Technol. 297, 122492. https://
(Tunçbilek) and effect of temperature and pressure on product distribution. Energy doi.org/10.1016/j.biortech.2019.122492.
Convers. Manag. 46 (13), 2185–2197. Jones, S.B., Valkenburt, C., Walton, C.W., Elliott, D.C., Holladay, J.E., Stevens, D.J.,
Carlson, T.R., Cheng, Y.-T., Jae, J., Huber, G.W., 2011. Production of green aromatics Kinchin, C., Czernik, S. 2009. Production of gasoline and diesel from biomass via fast
and olefins by catalytic fast pyrolysis of wood sawdust. Energy Environ. Sci. 4 (1), pyrolysis, hydrotreating and hydrocracking: a design case. Pacific Northwest
145–161. National Laboratory (PNNL).
Carpenter, D., Westover, T.L., Czernik, S., Jablonski, W., 2014. Biomass feedstocks for Jung, J.-M., Oh, J.-I., Baek, K., Lee, J., Kwon, E.E., 2018. Biodiesel production from waste
renewable fuel production: a review of the impacts of feedstock and pretreatment on cooking oil using biochar derived from chicken manure as a porous media and
the yield and product distribution of fast pyrolysis bio-oils and vapors. Green Chem. catalyst. Energy Convers. Manag. 165, 628–633.
16 (2), 384–406. Jung, S., Lee, J., Park, Y.-K., Kwon, E.E., 2020. Bioelectrochemical systems for a circular
Chandler, D.S., Resende, F.L.P., 2019. Comparison between catalytic fast pyrolysis and bioeconomy. Bioresour. Technol. 300, 122748.
catalytic fast hydropyrolysis for the production of liquid fuels in a fluidized bed Kawale, H.D., Kishore, N., 2019. Production of hydrocarbons from a green algae
reactor. Energy Fuel. 33 (4), 3199–3209. (Oscillatoria) with exploration of its fuel characteristics over different reaction
Chandler, D.S., Seufitelli, G.V.S., Resende, F.L.P., 2020. Catalytic route for the atmospheres. Energy 178, 344–355.
production of alkanes from hydropyrolysis of biomass. Energy Fuel. 34 (10), Khodier, A., Kilgallon, P., Legrave, N., Simms, N., Oakey, J., Bridgwater, T., 2009. Pilot-
12573–12585. scale combustion of fast-pyrolysis bio-oil: ash deposition and gaseous emissions.
Chang, Z., Duan, P., Xu, Y., 2015. Catalytic hydropyrolysis of microalgae: influence of Environ. Prog. Sustainable Energy 28 (3), 397–403.
operating variables on the formation and composition of bio-oil. Bioresour. Technol. Kim, S., Kwon, E.E., Kim, Y.T., Jung, S., Kim, H.J., Huber, G.W., Lee, J., 2019a. Recent
184, 349–354. advances in hydrodeoxygenation of biomass-derived oxygenates over heterogeneous
Chareonpanich, M., Boonfueng, T., Limtrakul, J., 2002. Production of aromatic catalysts. Green Chem. 21 (14), 3715–3743.
hydrocarbons from Mae-Moh lignite. Fuel Process. Technol. 79 (2), 171–179. Kim, S., Tsang, Y.F., Kwon, E.E., Lin, K.-Y.-A., Lee, J., 2019b. Recently developed
Che, Q., Yang, M., Wang, X., Yang, Q., Rose Williams, L., Yang, H., Zou, J., Zeng, K., methods to enhance stability of heterogeneous catalysts for conversion of biomass-
Zhu, Y., Chen, Y., Chen, H., 2019. Influence of physicochemical properties of metal derived feedstocks. Korean J. Chem. Eng. 36 (1), 1–11.
modified ZSM-5 catalyst on benzene, toluene and xylene production from biomass Kloekhorst, A., Shen, Y., Yie, Y., Fang, M., Heeres, H.J., 2015. Catalytic
catalytic pyrolysis. Bioresour. Technol. 278, 248–254. hydrodeoxygenation and hydrocracking of Alcell® lignin in alcohol/formic acid
Chen, L., Ma, X., Tang, F., Li, Y., Yu, Z., Chen, X., 2020a. Comparison of catalytic effect mixtures using a Ru/C catalyst. Biomass Bioenergy 80, 147–161.
on upgrading bio-oil derived from co-pyrolysis of water hyacinth and scrap tire over Klose, W., Stuke, V., 1993. Comparison of the pyrolysis of different types of biomass and
multilamellar MFI nanosheets and HZSM-5. Bioresour. Technol. 312, 123592. coals. Fuel Process. Technol. 36 (1), 283–289.
Chen, W., Lu, J., Zhang, C., Xie, Y., Wang, Y., Wang, J., Zhang, R., 2020b. Aromatic Kong, L., Zhang, L., Gu, J., Gou, L., Xie, L., Wang, Y., Dai, L., 2020. Catalytic
hydrocarbons production and synergistic effect of plastics and biomass via one-pot hydrotreatment of kraft lignin into aromatic alcohols over nickel-rhenium supported
catalytic co-hydropyrolysis on HZSM-5. J. Anal. Appl. Pyrolysis 147, 104800. on niobium oxide catalyst. Bioresour. Technol. 299, 122582.
Choudhary, P., Malik, A., Pant, K.K., 2020. Exploration of a novel biorefinery based on Kumar, R., Strezov, V., Lovell, E., Kan, T., Weldekidan, H., He, J., Dastjerdi, B., Scott, J.,
sequential hydropyrolysis and anaerobic digestion of algal biofilm: a comprehensive 2019. Bio-oil upgrading with catalytic pyrolysis of biomass using Copper/zeolite-
characterization of products for energy and chemical production. Sustain. Energy Nickel/zeolite and Copper-Nickel/zeolite catalysts. Bioresour. Technol. 279,
Fuels 4 (3), 1481–1495. 404–409.
Dabros, T.M.H., Stummann, M.Z., Høj, M., Jensen, P.A., Grunwaldt, J.-D., Gabrielsen, J., Kwon, E.E., Lee, T., Ok, Y.S., Tsang, D.C.W., Park, C., Lee, J., 2018. Effects of calcium
Mortensen, P.M., Jensen, A.D., 2018. Transportation fuels from biomass fast carbonate on pyrolysis of sewage sludge. Energy 153, 726–731.
pyrolysis, catalytic hydrodeoxygenation, and catalytic fast hydropyrolysis. Prog. Lappas, A.A., Bezergianni, S., Vasalos, I.A., 2009. Production of biofuels via co-
Energy Combust. Sci. 68, 268–309. processing in conventional refining processes. Catal. Today 145 (1), 55–62.
Dai, L., Wang, Y., Liu, Y., Ruan, R., Duan, D., Zhao, Y., Yu, Z., Jiang, L., 2019. Catalytic Lappas, A.A., Kalogiannis, K.G., Iliopoulou, E.F., Triantafyllidis, K.S., Stefanidis, S.D.,
fast pyrolysis of torrefied corn cob to aromatic hydrocarbons over Ni-modified 2012. Catalytic pyrolysis of biomass for transportation fuels. WIREs Energy Environ.
hierarchical ZSM-5 catalyst. Bioresour. Technol. 272, 407–414. 1 (3), 285–297.
Dayton, D.C., Carpenter, J., Farmer, J., Turk, B., Gupta, R., 2013. Biomass hydropyrolysis Lee, D.-J., Lu, J.-S., Chang, J.-S., 2020a. Pyrolysis synergy of municipal solid waste
in a pressurized fluidized bed reactor. Energy Fuel. 27 (7), 3778–3785. (MSW): a review. Bioresour. Technol. 318, 123912.
Dayton, D.C., Hlebak, J., Carpenter, J.R., Wang, K., Mante, O.D., Peters, J.E., 2016. Lee, J., Kwon, E.E., Park, Y.-K., 2020b. Recent advances in the catalytic pyrolysis of
Biomass hydropyrolysis in a fluidized bed reactor. Energy Fuel. 30 (6), 4879–4887. microalgae. Catal. Today 355, 263–271.
Foong, S.Y., Chan, Y.H., Cheah, W.Y., Kamaludin, N.H., Tengku Ibrahim, T.N.B., Lee, J., Sarmah, A.K., Kwon, E.E., 2019. Production and formation of biochar. In: Ok, Y.
Sonne, C., Peng, W., Show, P.-L., Lam, S.S., 2021. Progress in waste valorization S., Tsang, D.C.W., Bolan, N. (Eds.), Biochar from Biomass and Waste. Elsevier, J.M.
using advanced pyrolysis techniques for hydrogen and gaseous fuel production. Novak, pp. 3–18.
Bioresour. Technol. 320, 124299. Lee, X.J., Ong, H.C., Gan, Y.Y., Chen, W.-H., Mahlia, T.M.I., 2020c. State of art review on
Galiasso, R., González, Y., Lucena, M., 2014. New inverted cyclone reactor for flash conventional and advanced pyrolysis of macroalgae and microalgae for biochar, bio-
hydropyrolysis. Catal. Today 220–222, 186–197. oil and bio-syngas production. Energy Convers. Manag. 210, 112707.
Gamliel, D.P., Bollas, G.M., Valla, J.A., 2017a. Bifunctional Ni-ZSM-5 catalysts for the Leflaive, P., Lemberton, J.L., Pérot, G., Mirgain, C., Carriat, J.Y., Colin, J.M., 2002. On
pyrolysis and hydropyrolysis of biomass. Energy Technol. 5 (1), 172–182. the origin of sulfur impurities in fluid catalytic cracking gasoline—reactivity of
Gamliel, D.P., Bollas, G.M., Valla, J.A., 2018. Two-stage catalytic fast hydropyrolysis of thiophene derivatives and of their possible precursors under FCC conditions. Appl.
biomass for the production of drop-in biofuel. Fuel 216, 160–170. Catal. A: Gen. 227 (1-2), 201–215.
Gamliel, D.P., Wilcox, L., Valla, J.A., 2017b. The effects of catalyst properties on the Li, C., Aston, J.E., Lacey, J.A., Thompson, V.S., Thompson, D.N., 2016. Impact of
conversion of biomass via catalytic fast hydropyrolysis. Energy Fuel. 31 (1), feedstock quality and variation on biochemical and thermochemical conversion.
679–687. Renew. Sustain. Energ. Rev. 65, 525–536.
Hassan, H., Lim, J.K., Hameed, B.H., 2019. Catalytic co-pyrolysis of sugarcane bagasse Li, K., Zhang, G., Wang, Z.-X., Hu, B., Lu, Q., 2020. Calcium formate assisted catalytic
and waste high-density polyethylene over faujasite-type zeolite. Bioresour. Technol. pyrolysis of pine for enhanced production of monocyclic aromatic hydrocarbons
284, 406–414. over bimetal-modified HZSM-5. Bioresour. Technol. 315, 123805.
He, Y., Zhao, Y., Chai, M., Zhou, Z., Sarker, M., Li, C., Liu, R., Cai, J., Liu, X., 2020. Linck, M., Felix, L., Marker, T., Roberts, M., 2014. Integrated biomass hydropyrolysis and
Comparative study of fast pyrolysis, hydropyrolysis and catalytic hydropyrolysis of hydrotreating: a brief review. WIREs Energy Environ. 3 (6), 575–581.
poplar sawdust and rice husk in a modified Py-GC/MS microreactor system: insights Liu, C., Wang, H., Karim, A.M., Sun, J., Wang, Y., 2014. Catalytic fast pyrolysis of
into product distribution, quantum description and reaction mechanism. Renew. lignocellulosic biomass. Chem. Soc. Rev. 43 (22), 7594–7623.
Sust. Energ. Rev. 119, 109604. https://doi.org/10.1016/j.rser.2019.109604. Liu, Q., Bai, Y., Chen, H., Chen, M., Sang, Y., Wu, K., Ma, Z., Ma, Y., Li, Y., 2021a.
Hu, Y., Li, J., Wang, S., Xu, L., Barati, B., Cao, B., Wang, H., Xie, K., Wang, Q., 2021. Catalytic conversion of enzymatic hydrolysis lignin into cycloalkanes over a gamma-
Catalytic fast hydropyrolysis of seaweed biomass with different zeolite catalysts to alumina supported nickel molybdenum alloy catalyst. Bioresour. Technol. 323,
produce high-grade bio-oil. Process Saf. Environ. Prot. 146, 69–76. 124634.
Huang, Y., Qiu, S., Oduro, I.N., Guo, X., Fang, Y., 2016. Production of high-yield bio-oil Liu, Y., Wu, S., Zhang, H., Xiao, R., 2021b. Fast pyrolysis of torrefied holocellulose for
with a high effective hydrogen/carbon molar ratio through acidolysis and in situ producing long-chain ether precursors in a fluidized bed. Bioresour. Technol. 341,
hydrogenation. Energy Fuel. 30 (11), 9524–9531. 125770.
Huber, G.W., Iborra, S., Corma, A., 2006. Synthesis of transportation fuels from biomass: Love, G.D., Bowden, S.A., Jahnke, L.L., Snape, C.E., Campbell, C.N., Day, J.G.,
chemistry, catalysts, and engineering. Chem. Rev. 106 (9), 4044–4098. Summons, R.E., 2005. A catalytic hydropyrolysis method for the rapid screening of
Iliopoulou, E.F., Stefanidis, S.D., Kalogiannis, K.G., Delimitis, A., Lappas, A.A., microbial cultures for lipid biomarkers. Org. Geochem. 36 (1), 63–82.
Triantafyllidis, K.S., 2012. Catalytic upgrading of biomass pyrolysis vapors using Mahesh, D., Ahmad, S., Kumar, R., Chakravarthy, S.R., Vinu, R., 2021. Hydrothermal
transition metal-modified ZSM-5 zeolite. Appl. Catal. B: Environ. 127, 281–290. liquefaction of municipal solid wastes for high quality bio-crude production using
Jan, O., Marchand, R., Anjos, L.C.A., Seufitelli, G.V.S., Nikolla, E., Resende, F.L.P., 2015. glycerol as co-solvent. Bioresour. Technol. 339, 125537.
Hydropyrolysis of lignin using Pd/HZSM-5. Energy Fuel. 29 (3), 1793–1800.

9
S. Oh et al. Bioresource Technology 342 (2021) 126067

Mante, O.D., Dayton, D.C., Gabrielsen, J., Ammitzboll, N.L., Barbee, D., Verdier, S., brown coal–polyolefinic plastic co-pyrolysis behavior. J. Anal. Appl. Pyrolysis 78
Wang, K., 2016. Integration of catalytic fast pyrolysis and hydroprocessing: a (2), 257–264.
pathway to refinery intermediates and “drop-in” fuels from biomass. Green Chem. 18 Singh, N.R., Delgass, W.N., Ribeiro, F.H., Agrawal, R., 2010. Estimation of liquid fuel
(22), 6123–6135. yields from biomass. Environ. Sci. Technol. 44 (13), 5298–5305.
Marker, T.L., Felix, L.G., Linck, M.B., Roberts, M.J., 2012. Integrated hydropyrolysis and Sirous-Rezaei, P., Jae, J., Cho, K., Ko, C.H., Jung, S.-C., Park, Y.-K., 2019. Insight into the
hydroconversion (IH2) for the direct production of gasoline and diesel fuels or effect of metal and support for mild hydrodeoxygenation of lignin-derived phenolics
blending components from biomass, part 1: proof of principle testing. Environ. Prog. to BTX aromatics. Chem. Eng. J. 377, 120121.
Sustain. Energy 31 (2), 191–199. Sirous-Rezaei, Pouya, Jae, Jungho, Ha, Jeong-Myeong, Ko, Chang Hyun, Kim, Ji Man,
Marker, T.L., Felix, L.G., Linck, M.B., Roberts, M.J., Ortiz-Toral, P., Wangerow, J., 2014. Jeon, Jong-Ki, Park, Young-Kwon, 2018. Mild hydrodeoxygenation of phenolic
Integrated hydropyrolysis and hydroconversion (IH2®) for the direct production of lignin model compounds over a FeReOx/ZrO2 catalyst: zirconia and rhenium oxide
gasoline and diesel fuels or blending components from biomass, part 2: continuous as efficient dehydration promoters. Green Chem. 20 (7), 1472–1483.
testing. Environ. Prog. Sustain. Energy 33 (3), 762–768. Sirous-Rezaei, P., Park, Y.-K., 2020. Catalytic hydropyrolysis of lignin: suppression of
Meesuk, S., Cao, J.-P., Sato, K., Ogawa, Y., Takarada, T., 2011. Fast pyrolysis of rice husk coke formation in mild hydrodeoxygenation of lignin-derived phenolics. Chem. Eng.
in a fluidized bed: effects of the gas atmosphere and catalyst on bio-oil with a J. 386, 121348.
relatively low content of oxygen. Energy Fuel. 25 (9), 4113–4121. Stummann, M.Z., Elevera, E., Hansen, A.B., Hansen, L.P., Beato, P., Davidsen, B.,
Melligan, F., Hayes, M.H.B., Kwapinski, W., Leahy, J.J., 2012. Hydro-pyrolysis of Wiwel, P., Gabrielsen, J., Jensen, P.A., Jensen, A.D., Høj, M., 2020a. Catalytic
biomass and online catalytic vapor upgrading with Ni-ZSM-5 and Ni-MCM-41. hydropyrolysis of biomass using supported CoMo catalysts – effect of metal loading
Energy Fuel. 26 (10), 6080–6090. and support acidity. Fuel 264, 116807.
Melligan, F., Hayes, M.H.B., Kwapinski, W., Leahy, J.J., 2013. A study of hydrogen Stummann, M.Z., Høj, M., Davidsen, B., Hansen, A.B., Hansen, L.P., Wiwel, P.,
pressure during hydropyrolysis of Miscanthus x giganteus and online catalytic vapour Schandel, C.B., Gabrielsen, J., Jensen, P.A., Jensen, A.D., 2020b. Effect of the
upgrading with Ni on ZSM-5. J. Anal. Appl. Pyrolysis 103, 369–377. catalyst in fluid bed catalytic hydropyrolysis. Catal. Today 355, 96–109.
Mo, L., Dai, H., Feng, L.i., Liu, B., Li, X., Chen, Y., Khan, S., 2020. In-situ catalytic Stummann, M.Z., Høj, M., Davidsen, B., Hansen, L.P., Beato, P., Gabrielsen, J., Jensen, P.
pyrolysis upgradation of microalgae into hydrocarbon rich bio-oil: effects of nitrogen A., Jensen, A.D., 2019a. Deactivation of a CoMo catalyst during catalytic
and carbon dioxide environment. Bioresour. Technol. 314, 123758. https://doi.org/ hydropyrolysis of biomass. part 2: characterization of the spent catalysts and char.
10.1016/j.biortech.2020.123758. Energy. Fuel. 33 (12), 12387–12402.
Mohan, D., Pittman, C.U., Steele, P.H., 2006. Pyrolysis of wood/biomass for bio-oil: a Stummann, M.Z., Høj, M., Gabrielsen, J., Clausen, L.R., Jensen, P.A., Jensen, A.D., 2021.
critical review. Energy Fuel. 20 (3), 848–889. A perspective on catalytic hydropyrolysis of biomass. Renew. Sust. Energ. Rev. 143,
Moogi, S., Jae, J., Kannapu, H.P.R., Ahmed, A., Park, E.D., Park, Y.-K., 2020. 110960.
Enhancement of aromatics from catalytic pyrolysis of yellow poplar: role of Stummann, M.Z., Høj, M., Hansen, A.B., Beato, P., Wiwel, P., Gabrielsen, J., Jensen, P.A.,
hydrogen and methane decomposition. Bioresour. Technol. 315, 123835. Jensen, A.D., 2019b. Deactivation of a CoMo catalyst during catalytic hydropyrolysis
Morvay, Z., Gvozdenac, D., 2008. Applied Industrial Energy and Environmental of biomass. part 1. product distribution and composition. Energy. Fuel. 33 (12),
Management. John Wiley & Sons. 12374–12386.
Nguyen, T.-V., Clausen, L.R., 2019. Techno-economic analysis of polygeneration systems Stummann, M.Z., Høj, M., Hansen, A.B., Davidsen, B., Wiwel, P., Gabrielsen, J.,
based on catalytic hydropyrolysis for the production of bio-oil and fuels. Energy Jensen, P.A., Jensen, A.D., 2019c. New insights into the effect of pressure on
Convers. Manag. 184, 539–558. catalytic hydropyrolysis of biomass. Fuel Process. Technol. 193, 392–403.
Nguyen, T.S., Duong, T.L., Pham, T.T.T., Nguyen, D.T., Le, P.N., Nguyen, H.L., Huynh, T. Stummann, M.Z., Høj, M., Schandel, C.B., Hansen, A.B., Wiwel, P., Gabrielsen, J.,
M., 2017. Online catalytic deoxygenation of vapour from fast pyrolysis of Jensen, P.A., Jensen, A.D., 2018. Hydrogen assisted catalytic biomass pyrolysis.
Vietnamese sugarcane bagasse over sodium-based catalysts. J. Anal. Appl. Pyrolysis effect of temperature and pressure. Biomass Bioenergy 115, 97–107.
127, 436–443. Stummann, M.Z., Hansen, A.B., Hansen, L.P., Davidsen, B., Rasmussen, S.B., Wiwel, P.,
No, S.-Y., 2014. Application of bio-oils from lignocellulosic biomass to transportation, Gabrielsen, J., Jensen, P.A., Jensen, A.D., Høj, M., 2019d. Catalytic hydropyrolysis of
heat and power generation—a review. Renew. Sust. Energ. Rev. 40, 1108–1125. biomass using molybdenum sulfide based catalyst. effect of promoters. Energy. Fuel.
Onay, O., Gaines, A.F., Kockar, O.M., Adams, M., Tyagi, T.R., Snape, C.E., 2006. 33 (2), 1302–1313.
Comparison of the generation of oil by the extraction and the hydropyrolysis of Talmadge, M.S., Baldwin, R.M., Biddy, M.J., McCormick, R.L., Beckham, G.T.,
biomass. Fuel 85 (3), 382–392. Ferguson, G.A., Czernik, S., Magrini-Bair, K.A., Foust, T.D., Metelski, P.D.,
Perkins, G., Batalha, N., Kumar, A., Bhaskar, T., Konarova, M., 2019. Recent advances in Hetrick, C., Nimlos, M.R., 2014. A perspective on oxygenated species in the refinery
liquefaction technologies for production of liquid hydrocarbon fuels from biomass integration of pyrolysis oil. Green Chem. 16 (2), 407–453.
and carbonaceous wastes. Renew. Sust. Energ. Rev. 115, 109400. Tan, E.C.D., Marker, T.L., Roberts, M.J., 2014. Direct production of gasoline and diesel
Perkins, G., Bhaskar, T., Konarova, M., 2018. Process development status of fast pyrolysis fuels from biomass via integrated hydropyrolysis and hydroconversion process—a
technologies for the manufacture of renewable transport fuels from biomass. Renew. techno-economic analysis. Environ. Prog. Sustain. Energy 33 (2), 609–617.
Sust. Energ. Rev. 90, 292–315. Thangalazhy-Gopakumar, S., Adhikari, S., Gupta, R.B., 2012. Catalytic pyrolysis of
Pinho, A.d.R., de Almeida, M.B.B., Mendes, F.L., Ximenes, V.L., Casavechia, L.C. 2015. biomass over H+ZSM-5 under hydrogen pressure. Energy Fuel. 26 (8), 5300–5306.
Co-processing raw bio-oil and gasoil in an FCC Unit. Fuel Process. Technol., 131, Tian, Y., Li, J., Wei, W., Zong, P., Zhang, D., Zhu, Y., Qiao, Y., 2021. Parametric effect of
159-166. biomass partial hydropyrolysis process in a downer reactor to co-produce high-
Pinto, F., Miranda, M., Costa, P., 2016. Production of liquid hydrocarbons from rice crop quality tar and syngas. Bioresour. Technol. 320, 124401.
wastes mixtures by co-pyrolysis and co-hydropyrolysis. Fuel 174, 153–163. Valle, B., Gayubo, A.G., Aguayo, A.T., Olazar, M., Bilbao, J., 2010. Selective production
Pirbazari, S.M., Norouzi, Omid, Kohansal, Komeil, Tavasoli, Ahmad, 2019. Experimental of aromatics by crude bio-oil valorization with a nickel-modified HZSM-5 zeolite
studies on high-quality bio-oil production via pyrolysis of Azolla by the use of a three catalyst. Energy Fuel. 24 (3), 2060–2070.
metallic/modified pyrochar catalyst. Bioresour. Technol. 291, 121802. https://doi. Venderbosch, R.H., 2015. A critical view on catalytic pyrolysis of biomass.
org/10.1016/j.biortech.2019.121802. ChemSusChem 8 (8), 1306–1316.
Resasco, D.E., Wang, B., Crossley, S., 2016. Zeolite-catalysed C-C bond forming reactions Venkatakrishnan, V.K., Degenstein, J.C., Smeltz, A.D., Delgass, W.N., Agrawal, R.,
for biomass conversion to fuels and chemicals. Catal. Sci. Technol. 6 (8), 2543–2559. Ribeiro, F.H., 2014. High-pressure fast-pyrolysis, fast-hydropyrolysis and catalytic
Resende, F.L.P., 2016. Recent advances on fast hydropyrolysis of biomass. Catal. Today hydrodeoxygenation of cellulose: production of liquid fuel from biomass. Green
269, 148–155. Chem. 16 (2), 792–802.
Santana Junior, J.A., Menezes, A.L., Ataíde, C.H., 2019. Catalytic upgrading of fast Venkatakrishnan, V.K., Delgass, W.N., Ribeiro, F.H., Agrawal, R., 2015. Oxygen removal
hydropyrolysis vapors from industrial Kraft lignins using ZSM-5 zeolite and HY-340 from intact biomass to produce liquid fuel range hydrocarbons via fast-
niobic acid. J. Anal. Appl. Pyrolysis 144, 104720. hydropyrolysis and vapor-phase catalytic hydrodeoxygenation. Green Chem. 17 (1),
Saraeian, A., Nolte, M.W., Shanks, B.H., 2019. Deoxygenation of biomass pyrolysis 178–183.
vapors: Improving clarity on the fate of carbon. Renew. Sust. Energ. Rev. 104, Venkatesan, K., Prashanth, F., Kaushik, V., Choudhari, H., Mehta, D., Vinu, R., 2020.
262–280. Evaluation of pressure and temperature effects on hydropyrolysis of pine sawdust:
Shahabuddin, M., Krishna, Bhavya B., Bhaskar, Thallada, Perkins, Greg, 2020. Advances pyrolysate composition and kinetics studies. React. Chem. Eng. 5 (8), 1484–1500.
in the thermo-chemical production of hydrogen from biomass and residual wastes: Veses, A., Puértolas, B., Callén, M.S., García, T., 2015. Catalytic upgrading of biomass
Summary of recent techno-economic analyses. Bioresour. Technol. 299, 122557. derived pyrolysis vapors over metal-loaded ZSM-5 zeolites: effect of different metal
https://doi.org/10.1016/j.biortech.2019.122557. cations on the bio-oil final properties. Microporous Mesoporous Mater. 209,
Shang, J., Fu, G., Cai, Z., Feng, X., Tuo, Y., Zhou, X., Yan, H., Peng, C., Jin, X., Liu, Y., 189–196.
Chen, X., Yang, C., Chen, D., 2021. Regulating light olefins or aromatics production Wang, C., Lei, H., Zou, R., Qian, M., Mateo, W., Lin, X., Ruan, R., 2021. Biochar-driven
in ex-situ catalytic pyrolysis of biomass by engineering the structure of tin modified simplification of the compositions of cellulose-pyrolysis-derived biocrude oil coupled
ZSM-5 catalyst. Bioresour. Technol. 330, 124975. with the promotion of hydrogen generation. Bioresour. Technol. 334, 125251.
Sharma, S., Ghoshal, A.K., 2010. Study of kinetics of co-pyrolysis of coal and waste LDPE Wang, H., Male, J., Wang, Y., 2013. Recent advances in hydrotreating of pyrolysis bio-oil
blends under argon atmosphere. Fuel 89 (12), 3943–3951. and its oxygen-containing model compounds. ACS Catal. 3 (5), 1047–1070.
Sharypov, V.I., Beregovtsova, N.G., Kuznetsov, B.N., Baryshnikov, S.V., Cebolla, V.L., Wang, Jing-Xian, Cao, Jing-Pei, Zhao, Xiao-Yan, Liu, Sheng-Nan, Ren, Xue-Yu,
Weber, J.V., Collura, S., Finqueneisel, G., Zimny, T., 2006. Co-pyrolysis of wood Zhao, Ming, Cui, Xin, Chen, Qiang, Wei, Xian-Yong, 2019. Enhancement of light
biomass and synthetic polymers mixtures: Part IV: Catalytic pyrolysis of pine wood aromatics from catalytic fast pyrolysis of cellulose over bifunctional hierarchical
and polyolefinic polymers mixtures in hydrogen atmosphere. J. Anal. Appl. Pyrolysis HZSM-5 modified by hydrogen fluoride and nickel/hydrogen fluoride. Bioresour.
76 (1), 265–270. Technol. 278, 116–123.
Sharypov, V.I., Beregovtsova, N.G., Kuznetsov, B.N., Cebolla, V.L., Collura, S., Wright, M.M., Daugaard, D.E., Satrio, J.A., Brown, R.C., 2010. Techno-economic analysis
Finqueneisel, G., Zimny, T., Weber, J.V., 2007. Influence of reaction parameters on of biomass fast pyrolysis to transportation fuels. Fuel 89, S2–S10.

10
S. Oh et al. Bioresource Technology 342 (2021) 126067

Wu, Qiuhao, Wang, Yunpu, Jiang, Lin, Yang, Qi, Ke, Linyao, Peng, Yujie, Yang, Sha, Yasuda, H., Yamada, O., Kaiho, M., Nakagome, H., 2014. Effect of polyethylene addition
Dai, Leilei, Liu, Yuhuan, Ruan, Roger, 2020. Microwave-assisted catalytic upgrading to coal on hydrogasification enhancement. J. Mater. Cycles Waste Manag. 16 (1),
of co-pyrolysis vapor using HZSM-5 and MCM-41 for bio-oil production: co-feeding 151–155.
of soapstock and straw in a downdraft reactor. Bioresour. Technol. 299, 122611. Zhang, Jie, Zheng, Nan, Wang, Jie, 2018. Comparative investigation of rice husk,
https://doi.org/10.1016/j.biortech.2019.122611. thermoplastic bituminous coal and their blends in production of value-added
Xia, Q., Chen, Z., Shao, Y., Gong, X., Wang, H., Liu, X., Parker, S.F., Han, X., Yang, S., gaseous and liquid products during hydropyrolysis/co-hydropyrolysis. Bioresour.
Wang, Y., 2016. Direct hydrodeoxygenation of raw woody biomass into liquid Technol. 268, 445–453.
alkanes. Nat. Commun. 7 (1), 11162. Zhang, L., Gong, K., Lai, J., Alvey, P., 2017. Chemical composition and stability of
Xie, L.-F., Wang, F., Zhai, L.-L., Xu, Y.-P., Duan, P.-G., 2019. A sequenced batch process renewable hydrocarbon products generated from a hydropyrolysis vapor upgrading
for integrated hydropyrolysis and hydrotreatment of a microalgae and used engine process. Green Chem. 19 (15), 3628–3641.
oil blend. Fuel Process. Technol. 190, 47–54. Zhang, X., Yuan, Z., Yao, Q., Zhang, Y., Fu, Y., 2019. Catalytic fast pyrolysis of corn cob
Xu, Lujiang, He, Zijian, Zhang, Huan, Wu, Shenghong, Dong, Chengyu, Fang, Zhen, 2021. in ammonia with Ga/HZSM-5 catalyst for selective production of acetonitrile.
Production of aromatic amines via catalytic co-pyrolysis of lignin and phenol- Bioresour. Technol. 290, 121800.
formaldehyde resins with ammonia over commercial HZSM-5 zeolites. Bioresour. Zupko, R., 2019. Life cycle assessment of the production of gasoline and diesel from
Technol. 320, 124252. https://doi.org/10.1016/j.biortech.2020.124252. forest residues using integrated hydropyrolysis and hydroconversion. Int. J. Life
Xue, Y., Sharma, A., Huo, J., Qu, W., Bai, X., 2020. Low-pressure two-stage catalytic Cycle Assess. 24 (10), 1793–1804.
hydropyrolysis of lignin and lignin-derived phenolic monomers using zeolite-based
bifunctional catalysts. J. Anal. Appl. Pyrolysis 146, 104779.

11

You might also like