You are on page 1of 9

Fuel 276 (2020) 118112

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Effect of temperature on oil quality obtained through pyrolysis of sugarcane T


bagasse

Sascha Stegena, , Prasad Kaparajua
a
School of Engineering and Built Environment, Griffith University, Building N79, Nathan Campus, Nathan, 4111, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: The effect of temperature (300–650 °C) on the bio-oil quality and yields during pyrolysis of sugarcane bagasse
Bio-char was investigated. Results showed that temperature had a profound influence on the pyrolysis products’ yields
Biofuels and bio-oil quality. Elemental analyses of bio-char and bio-oil showed that the carbon content in bio-char in-
Bio-oil creased from 15.25% to 19.31% with increase in temperature from 300 °C to 500 °C and then decreased to
Fast pyrolysis
19.18% at 600 °C. Bio-oil and bio-char yields were seen to increase and decrease with an increase in pyrolysis
Sugarcane bagasse
temperature, respectively. Highest bio-oil yield (56.6% wt) was achieved at 500 °C while maximum amounts of
Temperature
bio-char (53.5% wt) and gas (32.9% wt) were obtained at 300 °C followed by at 400 °C. Highest HHV for bio-
char (11.78 MJ/kg) was recorded at 500 °C, which decreased with further increase in temperature to 600 °C.
Conversely, highest HHV for bio-oil (25.91 MJ/kg) was obtained at 300 °C. The obtained bio-oils were enriched
in lignin derivatives and sugars oligomers and their concentration increased with increase in temperatures from
450 to 550 °C. FTIR spectra revealed the presence of aromatics, ethers, alkanes, alkenes and other alcohols in
bio-char. Whilst alcohols, alkenes, ethers and other alcohols were noticed in the bio-oil. The complete absence of
aromatics in bio-oil facilitates long-term stability of the bio-oil. These results suggest that good yields and high-
quality bio-oil from sugarcane bagasse could be obtained at pyrolysis temperature of 550 °C. However, for an
efficient pyrolysis process with relatively high bio-oil yields, pyrolysis reactor should be operated at 450 °C.

1. Introduction: of people in Queensland, Australia [5]. However, the collection and


disposal of the sugarcane industry wastes poses significant challenge to
The major global concern on the depletion of fossil fuels and the the sustainable growth of the industry [5]. Previous studies have shown
increase in greenhouse gas emissions have given rise for research in that agricultural waste may increase soil hydrophobicity and decrease
alternative and clean fuels. Among the various renewable energy re- water retention and infiltration rates, which creates dead soil [6].
sources, biomass is considered as a unique resource as it provides CO2- Therefore, there is a shift in many oil producing countries towards more
neutral renewable energy and surpasses many other renewable energy sustainable methods of disposal [7].
sources in terms of its abundance, energy value and versatility. While In Australia, approximately 11 million tons of sugarcane bagasse is
bioenergy accounts for almost 18% in the global energy mix, in produced annually [8]. Traditionally, the produced bagasse is used for
Australia, 67 million tonnes of biomass and organic waste are generated onsite heat and electricity generation by most large and medium sized
every year [1]. Use of biomass as potential feedstock for bio-energy sugar mills [9]. Renewable energy production in Australia is projected
production has significant environmental and economic benefits [2]. It to increase to 9% by 2030 [10]. However, the Department of En-
not only reduces Australia’s dependence on fossil fuel and carbon vironment and Energy has set a target for large-scale generation of
footprint but can also contribute to greening the economy by creating 33,000 GWh in 2020. This means that about 23.5% of Australia's
jobs in clean technologies [3]. However, the use of biomass resources electricity generation in 2020 will be from renewable source.
for energy production is still under-utilised [4]. Furthermore, the dis- Decentralised energy generation is considered to solve many pro-
posal of these products creates substantial environmental and economic blems associated with the increasing in energy demand and costs. In
burdens. order to adopt new technologies, it must be affordable, easy and safe to
Sugar industry is the second largest agar-industry in Australia with operate. Among the biomass-to-bioenergy conversion technologies,
an annual revenue of $1.2 billion in export sales and employs thousands various potential thermo-chemical technologies were found to harvest


Corresponding author.
E-mail address: s.stegen@griffith.edu.au (S. Stegen).

https://doi.org/10.1016/j.fuel.2020.118112
Received 16 December 2019; Received in revised form 28 April 2020; Accepted 14 May 2020
0016-2361/ © 2020 Elsevier Ltd. All rights reserved.
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

energy from sugarcane bagasse, which include combustion, pyrolysis Table 1


and gasification [11]. These processes are mainly used to convert bio- Reactor parameters of maximum temperature, heating rate and calculated time
mass into energy-dense bio-fuels, namely bio-char, bio-oil and non- until maximum temperature is reached.
condensable gasses [12]. Depending upon the process conditions and Temperature [°C] Heating rate [K/s] Time until max temperature [s]
feedstocks used, thermo-chemical decomposition may produce different
products and ratios. 300 10 33.00
300 20 16.50
Pyrolysis is defined as the thermo-chemical decomposition of or-
300 30 11.00
ganic material to produce an array of solid, liquid and gaseous products 300 40 8.25
[13]. Thus, it is an attractive option for managing organic waste since it 400 10 38.00
results in significant waste volume reduction and the production of 400 20 19.00
400 30 12.67
value-added products such as bio-char, bio-oil and synthetic gas. The
400 40 9.50
yield of each product however varies, and is dependent on the pyrolysis 450 10 43.00
method, the feedstock properties, the heating rate, the final tempera- 450 20 21.50
ture, particle size as well as the catalytic agent used [14]. Depending 450 30 14.33
upon the process conditions, pyrolysis is classified as fast/flash, slow 450 40 10.75
500 10 48.00
and vacuum pyrolysis [15]. Fast pyrolysis is optimized for high bio-oil
500 20 24.00
yield, while vacuum pyrolysis is optimised for more even spread of solid 500 30 16.00
bio-char, bio-oil and syngas. On the contrary, slow pyrolysis is opti- 500 40 12.00
mised to yield bio-char as the main product [16]. The bio-char can be 550 10 53.00
550 20 26.50
upgraded to a high-value product called activated carbon, and is also
550 30 17.67
used by the sugar industry to clarify raw sugar for white sugar pro- 550 40 13.25
duction. 600 10 58.00
Bio-oil, also known as bio-crude or pyrolysis oil, is an energy pro- 600 20 29.00
duct of the pyrolysis process that can be easily stored for short time 600 30 19.33
600 40 14.50
periods, transported and handled. It can be used in the same way as
crude oil and be upgraded to be used as a substitute of heavy fuel oil
(bunker oil) or as a substitute to lower the crude oil consumption [17].
Table 2
It can also be upgraded using catalytic conversion to produce lighter
Chemical composition of the sugarcane bagasse.
transportation liquid fuels [18]. In fact, fast pyrolysis process can
convert up to 75% of the feedstock to a bio-oil intermediate suitable as Parameter Sugarcane bagasse

motor fuel [19]. Other uses of the product include pharmaceutical and Dry matter (%) 76.4
cosmetic production. Most of the research on bio-oil production Organic matter (dry wt, %) 84.4
through pyrolysis process was focused on the yield of the bio-oil and/or Moisture content (%) 23.6
syngas, especially by varying the pyrolysis temperature from 400 °C to Ash content (dry wt, %) 15.6
C (g/kg) 403.5
650 °C [20]. However, the energy density of the bio-oil is affected by
N (g/kg) 2.6
several factors such as the parent biomass characteristics and particle Glucan (dry wt, %) 37.5
size, the final temperature, heating rate, etc. [14]. Life cycle analyses Xylan (dry wt, %) 17.9
showed that optimising the yield does not necessarily result in optimum Galactan (dry wt, %) 1.5
energy returns [21]. Research into efficient energy content of the bio- Arabinan (dry wt, %) 2.5
Lignin (dry wt, %) 30.2
oil which would lead to a useful fuel for modern transportation road
vehicles is therefore needed. The overall efficiency should consider the
overall energy balance of the system using life cycle approach. How- concept and included into this chapter. Initial experiments showed
ever, no or limited literature was available on the impact of tempera- however same results by using the temperature range proposed in this
ture on the quality of bio-oil. This research is thus focused on the in- research project and thus, for visibility reasons, a heat resistant glass
vestigating the effect of temperature (300–600 °C) on the bio-oil quality container will be used for the execution of further experiments in re-
obtained during the pyrolysis of sugarcane bagasse. To achieve this, the gard to investigate the best temperature for the process.
research follows a practical approach and investigates the optimum
process condition in terms of temperature and residence time to max-
2.2. Reactor design
imise the bio-oil quality in a simple fast pyrolysis process.

The following equations were used to design the pyrolysis reactor


2. Materials and methods
and the amount of feedstock that can be processed as well as to cal-
culate the reaction time [22,23].
2.1. Feedstock and pyrolysis test rig
(a) The steam was determined by:
Sugarcane bagasse, which was provided by the Sunshine Sugar Mill ∑i 12Wp αi Ci
(Broadwater, New South Wales, Australia), was stored at room tem- Y=
Wf Fc (1)
perature (21 °C) throughout the experimental period. Prior to use, su-
garcane bagasse was dried for 60 days in ambient condition. The ma- where, Wp is the product per unit of time, ai is the number of carbon
terial was not milled and contained various particle sizes. The chemical atoms in the component represented by the species i, Ci is the mole
composition of sugarcane bagasse is presented in Table 2. fraction of i in the steam product, Wf is the feed inside the reactor in kg/
In this study, stainless steel fixed bed reactor was used to facilitate s, and Fc is the carbon fraction. The carbon molecular weight is 12 kg/
quick and uniform heat distribution within the furnace and especially kmol. The reaction rate was determined by the residence time τ:
with small amounts of feedstock. In addition, a fixed bed reactor is the
ln(1 − X c)
most cost-effective reactor. For the reactor, material must be chosen kg =
τ (2)
which does not react with the feedstock, vapour or heat. For this ex-
periment, a stainless-steel tube is constructed for the introduction of a (b) Carbon fraction Xc in the vapour stream was calculated by the

2
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

final carbon fraction C and the initial C0: (Ftot ) T P0


υ = υ0
(Ftot )0 T0 P (16)
C
Xc = 1 −
C0 (3) Based on the above calculations, the following process conditions
were used for the study. Feedstock particle sizes was between 0.2 and
(c) Chemical equations inside the reactor is assumed to be homo-
1 mm, while the moisture content was 15%. The reactor pressure was
genous due to good heat conductivity, and short connections between
calculated to vary from 1.6 atm for 300 °C to 3.3 atm for the highest
the reactor and the condensation train. The calculation can be pro-
temperature of 600 °C. The desired temperatures of 300 to 600 °C were
cessed with a stoichiometric equation with the species j over the re-
selected as previous studies have shown to report highest bio-oil out-
action:
puts between 450 °C and 550 °C [24]. At temperatures higher than
dNj 550 °C, secondary reactions of volatiles led to a significant lower bio-oil
Fjn + Gj = Fjn +
dt (4) yield. However, quality of bio-oil was reported to be better because of
the lower residence times, which was shown with various feedstocks
where, Gj is the rate of the species j formed inside the reactor.
including sugarcane bagasse [25]. The maximum temperature and
Furthermore, we can say that:
heating rates are calculated and can be seen in Table 1.
Gj = ∫ dGj = ∫ rj dV
VR VR (5) 2.3. Reactor operation

(d) For a packed bed reactor, we can add the following volume time
The pyrolysis reactor was consisting of a 150 ml glass Erlenmeyer
dependency for rj:
flask fitted with 2 or 3 holed glass stoppers. The reactor was connected
1 dNj to a condensation train and a vacuum system. As the initial heating
rj = (r)j V =
V dt (6) source was insufficient, flame heating was used as heat source for the
th rest of the experiments. The pipes leading from the reactor to the
(e) The design equation for the m independent reaction was cal-
condensation train were enough long to limit condensation before the
culated as:
traps. The condensation system consists of running tap water, which
np
dZ m t removed the organic vapours and gas products from the reactor through

= rm + ∑ (αkm rk) ⎛ ccr ⎞
⎜ ⎟
the condensation train. The condensable gases were then condensed in
k ⎝ 0 ⎠ (7)
the traps and recovered as liquid in a glass beaker (100 ml). The pro-
where, C0 is the concentration of the selected reference stream Ftot and: duced liquid was later weighed and analysed for chemical constituents.
In the pre-testing runs, it was shown that the fastest heating ramp was
Xm
Zm = achieved with the lowest residual times for the bio-oil within the re-
(Ftot)0 (8)
actor (data not shown). Therefore, the pyrolysis reactor was heated as
(f) With the reaction time tcr also freely selected: per the designed heating rate of 40 K/sec, residence time of 20 min and
final pyrolysis temperature (300, 400, 450, 500, 550 and 600 °C). The
Vr t sp
τ= = temperature was measured with a conventional thermometer in a glass
υ0 t cr t cr (9) body, to avoid contamination. The best temperature range was in-
(g) With the molar flow rate j and the volumetric flow rate υ, the vestigated in the second part of the experiment, when the reactor was
concentration of the conveniently selected reference stream can be operated at 300 °C to 600 °C. Each process condition was repeated
defined by: several times and two runs were carried out for each studied tem-
perature. In the end, 4 temperatures were selected and ranged from 400
j to 550 °C. Upon completion of experimental run, the reactor and the
Cj =
υ (10) condensation train were cooled under atmospheric pressure until the
(h) When the input stream is equal the reference stream: sample temperature was below 120 °C. Furthermore, the bio-oil con-
densation in the condensation train was washed out by using ethanol
nj nj
(Ftot )0 ⎛ solvent, added to the bio-oil sample and the bio-oil was then stripped
Cj =
υ ⎜
υj0 + ∑ (sj)m Zm⎞⎟ = C0 ⎛⎜υj0 + ∑ (sj)m Zm⎞⎟ from the solvent prior to analysis. The sample holder was then removed
⎝ m ⎠ ⎝ m ⎠ (11)
and weighed, after which the residue (bio-char) was removed and
(i) It can be assumed, that the density is almost constant: stored for further chemical analysis. A typical run would take between
υin = υ (12) 45 and 60 min, after which the reactor was cooled for 1 h depending on
the pyrolysis temperature employed. Prior to each run, the system was
(j) For the ith chemical reaction: flushed with nitrogen (99.9% pure) under atmospheric pressure to re-
γ (θ − 1)
'
move oxygen. The experiment then was carried out in a controlled
ri = ki (T0 )e θ ∙hi C j s (13) environment in fume hood.
where, hi C'j s
the function of species concentration, γ is the activation
energy and ki (T0) can be taken as the reaction rate is assumed to be 2.4. Analyses
constant with sufficient feed stock, constant heat supply, and stable gas
concentration. Furthermore, we can conclude that: Dry matter, ash content and moisture content were determined as
per the Standard Methods [26]. pH was monitored by a pH meter
nj
(Ftot )0 ⎧ (Ftot ) (Metrohm pH meter). Fourier-transform infrared spectroscopy (FTIR)
Cj =
υ ⎨ (Ftot )0 jn
γ + ∑ (sj)m Zm⎫⎬ analyses of the bio-crude were performed on a Nicolet 870 Nexus FTIR
⎩ m ⎭ (14)
spectrometer equipped with a Smart Endurance single bounce diamond
where, energy γ is constant: attenuated total reflectance (ATR) accessory (Nicolet Instrument Corp.,
γjn = γj0 US) as described elsewhere [27]. The spectrometer incorporated a KBr
(15)
beam splitter and a deuterated triglycine sulphate room temperature
(j) For the gas phase reaction, which means that the flow rate F, detector. Spectra were collected from 4000 cm to 1 to 525 cm-1 using
temperature T and pressure P is molar: 64 scans at 4 cm-1 resolution and a mirror velocity of 0.6329 cm s-1.

3
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

The measurement time for each spectrum was ~60 s. Which then results into,
The organic/bio-oil phase samples were analysed by Gas chroma-
Mbagasse = Mgas + Mbio − oil + Mbiochar (18)
tography–mass spectrometry using an Agilent 6890 Series Gas
Chromatograph and a HP 5973 mass spectrometer detector, employing
helium as the carrier gas [28] using an Agilent 6890 Series Gas Chro- 2.6. Energy balance
matograph and a HP 5973 mass spectrometer detector, employing he-
lium as the carrier gas. The installed column was an Agilent HP-1, The energy balance for each experimental run and operating tem-
30 m × 0.25 mm × 0.25 μm. Samples were injected with a split ratio of perature was calculate as the energy in (Qin) and energy out (Qout). As,
10:1 into the injection port set at 230 °C. The temperature program N2 gas does not undergo any phase change or reaction during the
commenced at 90 °C held for 10 min then heated at a rate of 3 °C.min−1 process, N2 gas requires energy only for heating and cooling to the
to a temperature of 260 °C. Once the temperature was reached the design temperatures. On the other hand, biomass requires heating, and
column was then held for another 5 min. Compounds were identified by then undergoes a reaction resulting in phase change. The produced
means of the Wiley library-HP G1035A and NIST library of mass spectra pyrolysis products require cooling and condensation as well.
and subsets-HP G1033A (a criteria quality value > 80% was used).
Qout = Qin + Qgenerated − − Qconsumed − −Qaccumulated + Qlosses (19)
Inductively Coupled Plasma-Optical Emission Spectroscopy (ICP-OES)
for the detection and quantification of trace and heavy metals was The ratio between the weight of solids collected in the reactor (Ybio-
carried out as described in [29]. Biomass compositional (cellulose, char,wt.%), the weight of tarry phase (Ybio-oil, wt.%) collected from the
hemicelluloses, lignin and ash) analysis of the bagasse and bio-char condenser train, which is operated with a constant flow of chilled water
solid samples were analysed according to the standard methods de- at 15 °C in an air-conditioned environment at 23 °C and the weight of
veloped by National Renewable Energy Laboratory (NREL), USA [30]. (dry) bagasse (Ybagasse wt.%) introduced into the system. The response,
Elemental analyses such as carbon, hydrogen, nitrogen, sulphur and Ygas, was not considered in this study due to the large deviation due to
oxygen content of the sugarcane bagasse and bio-char were determined the presence of losses.
by using a Flash EA 1112 Organic Analyser (Thermo Scientific, US)
[27]. Briefly, Samples of bio-char and bagasse were dried overnight at 3. Results and discussion
105 °C to constant weight prior to analysis. A subsample of dried bio-
char or bagasse (2–4 mg) was encapsulated in a tin container for the 3.1. Effect of pyrolysis temperature (300 °C to 600 °C) on chemical
measurement of carbon, hydrogen and nitrogen while another sub- composition of sugarcane bagasse
sample was encapsulated in a silver container for analysis of oxygen
content. The higher heating value (HHV) of the bagasse and bio-char Table 2 presents the major chemical and physical composition of the
samples were calculated by the Hydrogen, Carbon and Oxygen content sugarcane bagasse used. The feedstock sugarcane bagasse had a particle
including correction factors, based on the following equation [31]: size between 0.2 and 1 mm and contained 23.6% of moisture. The
moisture content was then lowered to 15% under room temperature
MJ ⎞
HHV ⎛⎜ ⎟ prior to the experimental runs. Elemental analysis showed that the
⎝ kg ⎠ composition of the sugarcane bagasse used in the present experiment is
= −1.3675 + 0.3137xCarbon + 0.7009xHydrogen + 0.0318xOxygen close to general elemental composition of sugarcane bagasse reported
(17) in the literature. However, the ash content in the present study feed-
stock is 15.6 wt% and is higher than the value reported in the literature
[34]. As ash contains the inorganic components e.g. metal oxides,
2.5. Mass balance pyrolysis temperature and residence time may lead to different yields of
products and contaminate products. The amount and/or distribution of
During pyrolysis, biomass undergoes a process that yields gas, liquid pyrolysis products viz., bio-char, bio-oil and gas are heavily dependent
and solid products. The product fractions were estimated from litera- on the initial composition of biomass. In Table 2, cellulose, hemi-cel-
ture to be 0.65, 0.2 and 0.15 for bio-oil, bio-char and syngas, respec- lulose and lignin fractions for sugarcane bagasse are presented. Results
tively [32,33]. Nitrogen is inert and is not produced by the pyrolysis showed that the glucan content in sugarcane bagasse was 37.5% (on
reaction. Therefore, inlet and outlet mass flow rates of nitrogen are dry weight basis) and is twice the concentration of xylan. Lignin content
equal. Bagasse mass (Mbagasse) + Mass of nitrogen was 30.2% (on dry weight basis). The higher heating value (HHV) of
(Mnitrogen) = Mnitrogen + Mass of syngas (Msyngas) + Mass of bio-oil raw sugarcane bagasse was determined to be 17 MJ/kg.
(Mbio-oil) + Mass of bio-char (Mbio-char) as N2 IN = N2 OUT. The elemental composition of pyrolysis products viz., bio-char and

Table 3
Elemental composition, C/H, C/O and calorific heating value (mf basis) of bio-char and bio-oil obtained during pyrolysis of sugarcane at 300–600 °C.
Bio-char

Max. Temp Nitrogen (%) Carbon (%) Hydrogen (%) Sulphur (%) Oxygen (%) C/H molar ratio C/O molar ratio HHV (MJ/kg)

300 °C 0.00 15.25 8.91 0.00 58.72 1.71 0.26 9.19


400 °C 0.00 19.08 9.06 0.00 53.49 2.11 0.37 11.29
500 °C 0.00 19.31 9.60 0.00 55.48 2.01 0.34 11.78
600 °C 0.00 19.18 9.16 0.00 69.46 2.09 0.27 9.67

Bio-oil

Max. Temp Nitrogen (%) Carbon (%) Hydrogen (%) Sulphur (%) Oxygen (%) C/H molar ratio C/O molar ratio HHV (MJ/kg)

300 °C 0.00 65.79 3.45 0.00 11.06 19.07 5.95 25.91


400 °C 0.52 56.53 3.06 0.00 11.59 18.47 4.88 22.47
500 °C 0.54 63.06 3.30 0.00 12.04 19.11 5.24 25.01
600 °C 0.50 59.95 2.53 0.00 6.43 23.66 9.32 23.62

4
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

bio-oil obtained during the pyrolysis of sugarcane bagasse at 300, 400, from 300 to 600 °C. On the other hand, the increase in C/H ratio with
500 and 600 °C is presented in Table 3. As expected, the ultimate increase in pyrolysis temperature indicates that stable and high-quality
analyses showed that the carbon content in bio-char increased from bio-bio-oil could be obtained at 600 °C.
15.25% to 19.31% with increase in temperature from 300 °C to 500 °C The product yields of each pyrolysis experimental run are presented
and then decreased to 19.18% with further increase in pyrolysis tem- in Table 4. Results show that temperature had a profound effect on
perature to 600 °C. In a similar study [35], reported that the fixed pyrolysis product yields and the quality of bio-oil at the tested tem-
carbon content during the fast pyrolysis of hemicellulose in sugarcane perature ranges (300 °C to 600 °C). Bio-char yields decreased with an
bagasse increased from 16.51% to 22.98% wt. with increase in tem- increase in temperature from 300 °C to 600 °C with the highest yield
perature from 550 to 850 °C. A similar trend was also noticed with obtained at 300 °C. The lower bio-char yields obtained at higher tem-
hydrogen and carbon to oxygen (C/O) molar ratio. However, C/H molar peratures is attributed to the fact that at higher temperature greater
ratio did not follow any trend. The possible reason for this discrepancy decomposition of organic material takes place promoting the release of
could be the generation of more thermostable compounds e.g. benzene volatile material [37]. By increasing the pyrolysis temperature from
at 600 °C through condensation and polymerisation of volatiles at 300 °C to 600 °C, a higher percentage of bio-oil and gas was obtained
higher temperature [35]. During the experimental run, volatile com- (Table 4). These results are in accordance with previous results in-
pounds were released and were absorbed in the condensation train; dicating that increased bio-char devolatilization at higher temperature
however, some of the residuals were entrapped in the bio-char. Pre- will results in a higher yield of liquid and gaseous products [37]. Both
vious studies have shown that volatile organic compounds such as these results suggest that the amount and type of volatile material
aromatic compounds, furfurals, anhydrosugars, organic acids and al- generated during the tested temperature range can result in the degree
cohols were predominately released as the temperatures during the of breakdown of cellulose, hemicellulose and lignin [38]. For instance,
pyrolysis of carbohydrates increase to 550 °C [36]. The decomposition a decrease in bio-char yield (from 26.8 to 10.7%) with a concomitant
products obtained during the pyrolysis of sugarcane bagasse at 300 to increase in gas yield (from 32 to 34.7%) was reported as the tem-
600 °C is presented in Table 6. With respect to the HHV, the highest perature was increased from 400 to 600 °C [39]. Previously, studies on
HHV of the bio-char was obtained at pyrolysis temperature of pyrolysis of lignocellulosic biomass [40–42] with main components,
400–500 °C (Table 4). On the other hand, the highest HHV for bio-oil i.e., hemicellulose, cellulose, and lignin showed that biomass pyrolysis
was obtained when the pyrolysis temperature was at 300 °C. However, can be considered as the sum of pyrolysis of the above three main
there was no clear trend in the composition of pyrolytic gas as the components [43]. However, thermal decomposition of these compo-
concentration and evolution rates of the main gas compounds (CO2, nents and evolution of volatile organic compounds is dependent upon
CO, H2, CH4, C2H4 and C2H6) changed with pyrolysis temperature the biomass composition, pyrolysis temperature, heating rate and mass
(Data not shown). After the initial start-up, CO2 was the dominant gas transfer [44]. Typical decomposition temperature ranges for cellulose
product at temperatures prior to 300 °C, followed by small amounts of and hemicelluloses are 240–390 °C and 160–360 °C, respectively [45].
CO and H2 and then light hydrocarbons. At 400 °C to 500 °C, both CO2 On the other hand, lignin decomposes at a slower rate but over a much
and CO were released with the maximum evolution occurring at wide temperature range of 180–900 °C [45]. For instance, thermo-
~400 °C. Both CO2 and CO accounted for more than 85% of the pyr- gravimetric analysis (TG) and its time derivative weight loss (DTG)
olysis gas volume. With increase in temperature from 500 to 600 °C, peaks during the pyrolysis of hemicellulose and cellulose in sugarcane
pyrolysis vapour was not effectively condensed leading to low pyrolysis bagasse were shown to be shifted to higher temperatures with in-
gas production (Table 4). On the other hand, hydrogen and carbon creasing heating rates, from 295 °C to 321 °C and from 350 °C to 376 °C,
concentration in bio-char were lowest at 300 °C whilst the oxygen level respectively [46]. At low heating rates and/or low temperature, there is
was the highest at 600 °C. This was expected due to the volatilisation of enough time for each individual component in the biomass to decom-
easily oxidised lower organic molecules at lower temperatures and pose. On the other hand, the decomposition of biomass component is
greater decomposition of higher organic material takes place as tem- almost simultaneous at high heating rates (Table 6) Thus, for full-scale
perature increases. This was evident from the carbon and hydrogen application, the amount of biomass and the size of the biomass is may
levels in bio-char and bio-oil. For bio-char, the carbon (15.25%) and affect the heat and transfer and may results the thermal decomposition
hydrogen (8.91%) content were lowest at 300 °C, whilst oxygen content behaviour and product yields.
was highest at 600 °C followed by at 300 °C. As expected, the ultimate A close observation of the FTIR spectra in Fig. 1, reveal a vast dif-
analyses showed that the carbon (65.8%) and hydrogen (3.45%) con- ference in the amount of aromatics produced during the pyrolysis
tents in bio-oil were highest at 300 °C followed by at 500 °C and 600 °C process. This was seen as an increase in the peak at the frequency of
(Table 3). However, highest C/H and C/O molar ratios in bio-oil were 1512 cm−1 and at 833 cm−1. These long molecule chains will cause the
noticed at 600 °C. The low C/H and C/O ratios noticed in bio-char potential fuel to be unstable and break down in a very short time [47].
samples, at all tested temperatures, than the untreated sugarcane ba- However, it can also be seen that there is a high content of ethers and
gasse indicates the preferential elimination of O and H relative to C in other alcohols as shown by vibrations at peak 1215 cm−1, which are
volatile matter. These results are in agreement with the results obtained promising as a potential fuel source. As the samples were not stripped of
by Crombie et al. [37]. In the above study, researchers have reported water, a peak at a frequency of 3335 cm−1 can be seen, which is not of
that small changes in H content in the bio-char had a proportionally concern once the fuel would be treated. Furthermore, a relatively high
larger effect on C/H molar ratio than the respective changes in O. Both amount of alkanes (1454 cm−1, 1378 cm−1, and 1362 cm−1) as well as
these ratios decreased in bio-char as the pyrolysis temperature rose alkenes (1605 cm−1, 1594 cm−1) can be seen. These compounds could
are reported to cause deposition of solids in combustion turbine as well
Table 4 as engines and thus consequently damage the machines permanently.
Mass balance average of sugarcane bagasse pyrolysis. Nevertheless, the relationship of alcohols and volatile substances is
shown.
Temperature Feedstock (g) Bio- Bio- Bio- Bio- Gas (g) Gas (%)
char char oil oil
(g) (%) (g) (%) 3.2. Effect of temperature range of 450 °C to 550 °C on bio-oil quality

300 °C 6.620 3.54 53.54 0.91 13.76 2.18 32.87 In the second phase of the experiment, pyrolysis temperatures were
400 °C 6.582 2.40 36.48 2.79 42.43 1.32 20.09
narrowed down to 450 °C, 500 °C, and 550 °C. The ultimate analyses, C/
500 °C 6.584 2.20 33.45 3.73 56.63 0.64 9.65
600 °C 6.622 2.26 34.15 3.47 52.39 0.88 13.33 H molar ratio and C/O molar ratio along with HHV values are presented
in Table 5. Prior to the analyses, samples were stripped off the H2O

5
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

Table 5
Elemental composition, C/H, C/O and calorific heating value (mf basis) of bio-oil obtained during pyrolysis of sugarcane at 450–550 °C.
Sample Nitrogen (%) Carbon (%) Hydrogen (%) Sulphur (%) Oxygen (%) C/H molar ratio C/O molar ratio HHV (MJ/kg)

450 °C 0.00 56.30 6.46 0.00 37.25 8.72 1.51 23.35


500 °C 0.00 56.08 6.40 0.00 37.53 8.76 1.49 24.85
550 °C 0.00 57.02 6.44 0.00 36.55 8.85 1.56 25.54

Table 6 noticed at vibrations 1232 cm−1 and 1159 cm−1. Similarly, alkenes
Composition analysis of solid output for narrowed down temperature range. were visible at wavelength 2966 cm−1, 2937 cm−1, 2849 cm−1,
Compounds Chemical structure 450 °C 500 °C 550 °C 1558 cm−1, and 1435 cm−1. Interestingly, aromatics were completely
(area (area (area %) absent in all samples indicting very high and long-term stability of the
%) %) bio-oil. Furthermore, no wax deposits, due to possible aromatics com-
pounds, were noticed even after the analysis was repeated four months
2,4-dimethyl phenol 3.20 0.54 2.41
after the experiments. Similar results were also obtained with respect to
the calculated HHV as well as the FTIR spectra of bio-oil.
The chemical composition of the bio-char at the three studied
4-ethyl phenol 25.42 18.78 19.83
temperatures is presented in Table 6. Results showed that temperature
had a profound effect on the type and amount of chemicals evolved at
2,3-dihydro-benzofuran 16.23 19.37 19.62
each studied pyrolysis temperature. This was evident in the difference
in evolution of different chemicals at the studied temperatures. High
4-ethyl-2-methoxy-phenol 5.70 3.29 6.03 concentration of mono-aromatic compounds such as phenols and its
derivatives, vanillin and benzoic acid were produced. However, the
concentration of the phenols and their derivatives increased with in-
2-methoxy-4-vinylphenol 7.97 5.69 8.41 crease in temperature from 450 °C to 600 °C. On the contrary, the
concentration of vanillin and benzoic acid decreased with increase in
temperature. This variation in the amount and the chemical composi-
2,6-dimethoxy-phenol 4.77 3.44 4.01
tion is dependent on the pyrolysis mechanism of lignin, which is con-
sidered to more complex compared to cellulose and hemicellulose. As
mentioned earlier, the pyrolysis of cellulose takes place by breakdown
Vanillin 3.01 0.00 0.00 of glycoside bonds at 350 °C. On the other hand, lignin degradation is
spread over a long range of temperatures and classified into primary
Apocynin 0.07 0.00 0.00
decomposition reactions at 200–400 °C and secondary reactions at
temperatures higher than 400 °C [48]. As the pyrolysis temperature was
raised from 450 °C to 550 °C, secondary pyrolysis reactions occurred
2,6-dimethoxy-4-(2- 2.61 4.49 4.21 transforming guaiacols/syringols to catechols/pyrogallols and o-cre-
propenyl)- phenol
sols/xylenols along with phenols. At 450–500 °C, methoxyl group-re-
lated reactions occur while in the temperature range of 500–550 °C,
Ethanol, 1-(4-hydroxy-3,5- 0.91 2.25 2.34 decomposition vanillin and apocynin occurred through gasification.
dimethoxyphenyl)- The absence of aromatics compounds such as vanillin and apocynin at
temperatures higher than 500 °C improves the stability of bio-oil and
Diethyl (4- 0.00 0.08 0.00 facilitate for longer storage times without risking the fuel quality and
biphenylylmethyl) forming solid deposits. Furthermore, the absence of aromatic compo-
phosphonate nents at the optimised temperature range would even enable the use in
2- (octyloxylcarbonyl) 0.00 0.17 0.15 modern engines and turbines. Previous studies have shown that aro-
benzoic acid
matic methoxy groups, which are stable during the primary decom-
position stage of pyrolysis, become reactive in the temperature range of
400–450 °C [49]. Therefore, the aromatic compounds that are produced
are mainly 4-substituted guaiacols from G-lignins and 4-substituted
content. Pyrolysis temperature had profound influence on the HHV syringols from S-lignins. On the other hand, the main volatile products
values and the values obtained in the second phase of the experiment from G-lignins included alcohol, aldehyde, isoeugenol, 4-vinyl guaiacol,
were lower than those obtained at 500 °C in the preliminary experi- vanillin, acetovanillone, and dihydroconiferyl alcohol [50]. In a similar
ments (Table 5). HHV values increased with increase in pyrolysis study Patwardhan et al. [2011] [49] reported the formation of mono-
temperature. The highest HHV value of 25.54 MJ/kg was obtained at meric phenolic compounds with phenol, 2,6-dimethoxy phenol, 2-
550 °C and it was expected as the oxygen content was low with slightly methoxy-4-vinyl phenol, and 4-vinyl phenol, as the major primary
higher values for H and C than at 450 °C and 500 °C. Previous studies products. The liquid bio-oil was rich in dimeric and oligomeric com-
reported that high bio-oil yield were obtained between 450 °C and pounds due to re-oligomerization of the monomeric products formed
550 °C [24]. At temperatures higher than 550 °C, secondary reactions of through condensation.
volatiles led to a significant lower bio-oil yield. However, quality of bio- After each run, a mass balance was performed to determine the solid
oil was reported to be better because of the lower residence times, bio-char and liquid bio-bio-oil yields, the gas yield was calculated as a
which was shown with various feedstocks including sugarcane bagasse difference between the two products. Although most of the bio-oil was
[25]. directly collected from the bottom of condensation collection vessel, a
The FTIR spectrum for temperature 450–550 °C is presented in small fraction was collected before and after cleaning the glass vessel.
Fig. 2. Temperature had a profound influence on the production of al- Acetone-oil formed during the washing of the larger units was collected
cohols and alkenes. Very high content of ethers and other alcohols were and weighed after evaporation. Mass balance of the pyrolysis products

6
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

Fig. 1. FTIR spectra of bio-oil obtained during the pyrolysis of sugarcane bagasse at a) 300 °C, b) 400 °C, c) 500 °C, and d) 600 °C.

obtained at 450 °C to 550 °C is presented in Table 7. Bio-oil yields no variation in the percentage of bio-char yields, as well as no varia-
decreased with increase in pyrolysis temperature from 450 °C to 550 °C. tions in the bio-oil content for these experimental conditions. The best
At 550 °C, bio-oil yields were decreased by 6.4%, but at the same time bio-oil output and yield was obtained at 550 °C.
its HHV was increased by 11%. These results agree with the values are Input energy Qin and output energy of the bio-oil Qout were used to
comparable with the values found in the literature [51]. Moreover, the estimate the energy balance and the results are presented in Table 8.
long term storage stability of the bio-oil, which is dependent on the O Based on the energy balance, pyrolysis of sugarcane bagasse was op-
content, has improved (Table 3 and Table 5). Thus, less energy is re- timal at 450 °C. Although the bio-oil quality as well as yield obtained at
quired for downstream processing of the fuel. Preliminary results 550 °C was higher than at 450 °C, the input energy needed to for the
showed that a mass loading up to the maximum, for the reactor size, led process was much higher at 550 °C than at 450 °C. Thus, considering the

Fig. 2. FTIR spectra of bio-oil obtained during the pyrolysis of sugarcane bagasse: a) Reference curve from worst previous sample, b) 450 °C, c) 500 °C, and d) 550 °C.

7
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

Table 7
Mass balance average of sugarcane bagasse pyrolysis.
Temperature Feedstock (g) Bio-char (g) Bio-char (%) Bio-oil (g) Bio-oil (%) Gas (g) Gas (%)

450 °C 6.550 2.175 33.206 3.971 60.623 0.404 6.171


500 °C 6.610 2.181 32.995 3.869 58.539 0.560 8.465
550 °C 6.662 2.136 32.062 3.717 55.801 0.809 12.137

Table 8 Acknowledgement
Energy balance for the pyrolysis process at the studied temperatures based on
the bio-oil HHV. The authors would like to thank Dr C L. Moghaddam and Professor
Temperature Qin (J/s) Qout (J/s) Efficiency in % William Doherty, from the Centre for Tropical Crops and Bio com-
modities, Queensland University of Technology, Australia, for their
300 °C 61158.28 21391.30 34.98 support in chemical analyses and sharing their experience with fuel
400 °C 83237.08 57209.52 68.73
analysis.
450 °C 94276.48 84915.66 90.07
500 °C 105315.88 87843.62 83.41
550 °C 116355.28 85493.42 73.48 References
600 °C 127394.68 74259.86 58.29
[1] Environment and Communications References Committee, Commonwealth of
Australia The Senate 2018.
overall energy and mass balance, the optimal temperature to carryout [2] Deb U, Bhuyan N, Bhattacharya SS, Kataki R. Characterization of agro-waste and
weed biomass to assess their potential for bioenergy production. Int J Renewable
pyrolysis of sugarcane bagasse is 450 °C.
Energy Dev 2019;8(3).
In this study, the feasibility of obtaining bio-oils with high con- [3] Maraseni T, Reardon-Smith K. Meeting national emissions reduction obligations: a
centration of phenols and other mono-aromatic compounds showed case study of Australia. Energies 2019;12(3):438.
[4] (2019). Global bioenergy statistics 2019.
that low temperature pyrolysis process has great potential of making
[5] Meghana M, Shastri Y. Sustainable valorization of sugar industry waste: Status,
bio-oils for chemical and transport industry. However, no attempt was opportunities, and challenges. Bioresour Technol 2020:122929.
made to enrich the bio-oil. Based on the results, it can be suggested that [6] Jiménez-de-Santiago DE, Yagüe MR, Bosch-Serra ÀD. Soil water repellency after
the pyrolysis reactor has to be operated at the desired temperature of slurry fertilization in a dryland agricultural system. Catena 2019;174:536–45.
[7] Pergola M, et al. Composting: The way for a sustainable agriculture. Appl Soil Ecol
450–550 °C and heating rate of 40 K/s in order to obtained high quality 2018;123:744–50.
bio-oil. However, bio-oil and bio-char yield obtained during the in- [8] Valix M, Katyal S, Cheung W. Chemical torrefaction as an alternative to established
dustrial full-scale process may depend on the mass and heat transfer, thermal technology for stabilisation of sugar cane bagasse as fuel. Environ Technol
2017;38(13–14):1638–43.
which might be affected when contact between the sample and the hot [9] Dias MO, Lima DR, Mariano AP. “Techno-Economic Analysis of Cogeneration of
zone, especially for larger feedstock. Less contact between the feedstock Heat and Electricity and Second-Generation Ethanol Production from Sugarcane,”
and heating source may result in temperature gradients and incomplete in Advances in Sugarcane. Biorefinery: Elsevier 2018:197–212.
[10] (2019). Australian Energy Update 2019.
reactions. [11] Soares L, Rabelo C, Delforno T, Silva E, Varesche M. Experimental design and
syntrophic microbial pathways for biofuel production from sugarcane bagasse
under thermophilic condition. Renewable Energy 2019;140:852–61.
4. Conclusion [12] Sarkar JK, Wang Q. Different Pyrolysis Process Conditions of South Asian Waste
Coconut Shell and Characterization of Gas, Bio-Char, and Bio-Oil. Energies
The study showed that pyrolysis of sugarcane bagasse at low tem- 2020;13(8):1970.
[13] Nandagopal V, Maheswari V, Anbarasi J. Pyrolysis Electricity Generation and
perature range from 300 to 600 °C was feasible. The optimum tem-
Biomass. J Comput Theor Nanosci 2019;16(2):428–9.
perature of the pyrolysis for maximum bio-oil output and best HHV was [14] Mishra RK, Mohanty K. Pyrolysis of three waste biomass: Effect of biomass bed
obtained at 550 °C. It was noted that an increase of the HHV of up to thickness and distance between successive beds on pyrolytic products yield and
properties. Renewable Energy 2019;141:549–58.
11% was possible. Further, absence of aromatics compounds such as
[15] A. Nzihou, B. Stanmore, N. Lyczko, and D. P. Minh, “The catalytic effect of inherent
vanillin and apocynin at temperatures higher than 500 °C can improve and adsorbed metals on the fast/flash pyrolysis of biomass,” 2019.
the stability of bio-oil and facilitate for longer storage times without [16] Vieira FR, Luna CMR, Arce GL, Ávila I. Optimization of slow pyrolysis process
further post-treatment. However, energy balance showed that the pyr- parameters using a fixed bed reactor for biochar yield from rice husk. Biomass
Bioenergy 2020;132:105412.
olysis process at 450⁰C was the most efficient temperature to produce [17] Syamsiro M, Saptoadi H, Kismurtono M, Mufrodi Z, Yoshikawa K. Utilization of
bio-oil for sugarcane bagasse. Although the amount of aromatics com- waste polyethylene pyrolysis oil as partial substitute for diesel fuel in a DI diesel
pounds produced was still low, the bio-oil yield as well as the HHV engine. Int J Smart Grid Clean Energy 2019;8(1):38–47.
[18] Valle B, García-Gómez N, Remiro A, Gayubo AG, Bilbao J. Cost-effective upgrading
obtained at 450 °C were comparable to the best bio-oil yields obtained of biomass pyrolysis oil using activated dolomite as a basic catalyst. Fuel Process
at 550 °C. Thus, pyrolysis of sugarcane bagasse should be performed at Technol 2019;195:106142.
550 °C for high bio-oil yields while efficient pyrolysis process with re- [19] Park JY, Kim J-K, Oh C-H, Park J-W, Kwon EE. Production of bio-oil from fast
pyrolysis of biomass using a pilot-scale circulating fluidized bed reactor and its
latively high bio-oil yields was obtained at 450 °C. characterization. J Environ Manage 2019;234:138–44.
[20] Okolie JA, Nanda S, Dalai AK, Kozinski JA. Hydrothermal gasification of soybean
straw and flax straw for hydrogen-rich syngas production: Experimental and ther-
CRediT authorship contribution statement modynamic modeling. Energy Convers Manage 2020;208:112545.
[21] Lu HR, El Hanandeh A. Life cycle perspective of bio-oil and biochar production from
Sascha Stegen: Conceptualization, Methodology, Writing - original hardwood biomass; what is the optimum mix and what to do with it? J Cleaner Prod
2019;212:173–89.
draft, Investigation, Writing - review & editing, Methodology. Prasad
[22] Froment GF, Bischoff KB, De Wilde J. Chemical reactor analysis and design. New
Kaparaju: Supervision, Resources, Writing - review & editing. York: Wiley; 1990.
[23] Mann U. Principles of chemical reactor analysis and design: new tools for industrial
chemical reactor operations. 2 ed. Hoboken, NJ, USA: John Wiley & Sons; 2009.
Declaration of Competing Interest [24] M. R. Rover, “Analysis of sugars and phenolic compounds in biooil,” Doctor of
Philosophy, Mechanical Engineering, Iowa State University, Iowa, 2013. [Online].
Available: https://lib.dr.iastate.edu/etd/13012.
The authors declare that they have no known competing financial [25] W.T. Tsai M.K. Lee Y.M. Chang Fast pyrolysis of rice straw, sugarcane bagasse and
interests or personal relationships that could have appeared to influ- coconut shell in an induction-heating reactor Journal of Analytical and Applied
ence the work reported in this paper. Pyrolysis 76 1–2 2006 pp. 230–237, 6// 10.1016/j.jaap.2005.11.007.

8
S. Stegen and P. Kaparaju Fuel 276 (2020) 118112

[26] Rabelo SC, Carrere H, Maciel Filho R, Costa AC. Production of bioethanol, methane 2012;114:644–53.
and heat from sugarcane bagasse in a biorefinery concept. Bioresour Technol [39] Miao X, Wu Q. High yield bio-oil production from fast pyrolysis by metabolic
2011;102(17):7887–95. https://doi.org/10.1016/j.biortech.2011.05.081. controlling of Chlorella protothecoides. J Biotechnol 2004;110(1):85–93. https://
[27] Tana T, et al. Structural changes of sugar cane bagasse lignin during cellulosic doi.org/10.1016/j.jbiotec.2004.01.013.
ethanol production process. ACS Sustainable Chem Eng 2016;4(10):5483–94. [40] Pasangulapati V, Ramachandriya KD, Kumar A, Wilkins MR, Jones CL, Huhnke RL.
[28] da Cunha ME, et al. Analysis of fractions and bio-oil of sugar cane straw by one- Effects of cellulose, hemicellulose and lignin on thermochemical conversion char-
dimensional and two-dimensional gas chromatography with quadrupole mass acteristics of the selected biomass. Bioresour Technol 2012;114:663–9.
spectrometry (GC×GC/qMS). Microchem J 2013;110:113–9. https://doi.org/10. [41] Burhenne L, Messmer J, Aicher T, Laborie M-P. The effect of the biomass compo-
1016/j.microc.2013.03.004. nents lignin, cellulose and hemicellulose on TGA and fixed bed pyrolysis. J Anal
[29] Popov BB, Najman S, Hristova VK, Ahmad MA. Inductively Coupled Plasma-Optical Appl Pyrol 2013;101:177–84.
Emission Spectroscopy (ICP-OES) Approach for the Determination of Heavy Metals, [42] Stefanidis SD, Kalogiannis KG, Iliopoulou EF, Michailof CM, Pilavachi PA, Lappas
Metalloid and Trace Element in Soil and Vegetables. Indian Horticulture J (IHJ) AA. A study of lignocellulosic biomass pyrolysis via the pyrolysis of cellulose,
2014:2249–6823. hemicellulose and lignin. J Anal Appl Pyrol 2014;105:143–50.
[30] Sluiter A, et al. Determination of structural carbohydrates and lignin in biomass. [43] Meng A, Chen S, Zhou H, Long Y, Zhang Y, Li Q. Pyrolysis and simulation of typical
Lab Anal Proc 2008;1617:1–16. components in wastes with macro-TGA. Fuel 2015;157:1–8.
[31] Sheng C, Azevedo J. Estimating the higher heating value of biomass fuels from basic [44] Onsree T, Tippayawong N, Zheng A, Li H. Pyrolysis behavior and kinetics of corn
analysis data. Biomass Bioenergy 2005;28(5):499–507. residue pellets and eucalyptus wood chips in a macro thermogravimetric analyzer.
[32] Huang L, Gu M. Effects of biochar on container substrate properties and growth of Case Stud Thermal Eng 2018;12:546–56.
plants—A review. Horticulturae 2019;5(1):14. [45] Varhegyi G, Antal Jr MJ, Szekely T, Szabo P. Kinetics of the thermal decomposition
[33] Brassard P, Godbout S, Raghavan V, Palacios JH, Grenier M, Zegan D. The pro- of cellulose, hemicellulose, and sugarcane bagasse. Energy Fuels 1989;3(3):329–35.
duction of engineered biochars in a vertical auger pyrolysis reactor for carbon se- [46] Hugo TJ. Pyrolysis of sugarcane bagasse. Stellenbosch: University of Stellenbosch;
questration. Energies 2017;10(3):288. 2010.
[34] Pippo W, Luengo C, Alberteris L, Garcia del Pino G, Neto J. Energy Recovery from [47] Heoa HS, et al. Influence of operation variables on fast pyrolysis of Miscanthus
Sugarcane: Study of Heating Value Variations of Sugarcane-trash with Moisture sinensis var. purpurascens. Elsevier Bioresource Technology May
Content during the. Milling Season 2014. 2010;101(10):3672–7.
[35] Peng Y, Wu S. Fast pyrolysis characteristics of sugarcane bagasse hemicellulose. [48] Patwardhan PR, Brown RC, Shanks BH. Understanding the fast pyrolysis of lignin.
Cellul Chem Technol 2011;45(9):605. ChemSusChem 2011;4(11):1629–36.
[36] Räisänen U, Pitkänen I, Halttunen H, Hurtta M. Formation of the main degradation [49] Patwardhan PR, Brown RC, Shanks BH. Product distribution from the fast pyrolysis
compounds from arabinose, xylose, mannose and arabinitol during pyrolysis. J of hemicellulose. ChemSusChem 2011;4:636.
Therm Anal Calorim 2003;72(2):481–8. [50] Choi YS, et al. Pyrolysis reaction networks for lignin model compounds: unraveling
[37] Crombie K, Mašek O, Sohi SP, Brownsort P, Cross A. The effect of pyrolysis con- thermal deconstruction of β-O-4 and α-O-4 compounds. Green Chem
ditions on biochar stability as determined by three methods. GCB Bioenergy 2016;18(6):1762–73.
2013;5(2):122–31. [51] Wang X, Kersten SRA, Prins W, van Swaaij WPM. Biomass pyrolysis in a fluidized
[38] Enders A, Hanley K, Whitman T, Joseph S, Lehmann J. Characterization of biochars bed reactor. Part 2: experimental validation of model results. Indust Eng Chem Res
to evaluate recalcitrance and agronomic performance. Bioresour Technol 2005;44(23):8786–95. https://doi.org/10.1021/ie050486y.

You might also like