Sie sind auf Seite 1von 6

Review of biomass pyrolysis oil properties and upgrading research

Zhang Qi
a,b,
*
, Chang Jie
a
, Wang Tiejun
a
, Xu Ying
a
a
Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China
b
University of Science and Technology of China, Hefei 230027, PR China
Received 9 November 2005; accepted 10 May 2006
Available online 22 June 2006
Abstract
Biomass fast pyrolysis liquefaction has aroused great attention and interests both at home and abroad extensively in recent years. This
paper reviews the physicochemical properties and discusses the characteristics of the components and compositions of biomass pyrolysis
oil. Furthermore, the problems and focuses were summarized with some suggestions presented on upgrading and applications of bio-oil
in the decades.
2006 Elsevier Ltd. All rights reserved.
Keywords: Biomass; Pyrolysis; Bio-oil; Properties; Upgrading
1. Introduction
Considering the fact that energy consumption is increas-
ing and limited fossil fuels are nearly exhausted, with
increasing populations and economic developments,
renewable energy should be widely explored in order to
renovate the energy sources structure and keep sustainable
development safe. Biomass is clean for it has negligible
content of sulphur, nitrogen and ash, which give lower
emissions of SO
2
, NO
x
and soot than conventional fossil
fuels. Zero net emission of CO
2
can be achieved because
CO
2
released from biomass will be recycled into the plants
by photosynthesis quantitatively. The energy crisis and fuel
tension made biomass fast pyrolysis liquefaction a more
important area of research and development [13]. Pyroly-
sis has been supported with the highest amount of funds
provided by the European Union for liquid bio-fuel tech-
nologies [4]. Bio-oil from biomass fast pyrolysis is mainly
produced from biomass residues in the absence of air at
atmospheric pressure, a low temperature (450550 C),
high heating rate (10
3
10
4
K/s) and short gas residence
time to crack into short chain molecules and be cooled to
liquid rapidly. Fast pyrolysis, an eective biomass conver-
sion with high liquid yield, as much as 7080% and a high
ratio of fuel to feed, is regarded as one of the reasonable
and promising technologies to compete with and eventually
replace non-renewable fossil fuel resources [5]. Recent
research on the physicochemical properties of biomass
pyrolysis oil was reviewed and some recommendations
were put forward based on the upgrading status and appli-
cation demands for future improvements.
2. Bio-oil
The liquid product from biomass pyrolysis is known as
biomass pyrolysis oil, and bio-oil, pyrolysis oil, or bio-crude
for short. Bio-oil is not a product of thermodynamic equi-
librium during pyrolysis but is produced with short reactor
times and rapid cooling or quenching from the pyrolysis
temperatures. This produces a condensate that is also not
at thermodynamic equilibrium at storage temperatures.
The chemical composition of the bio-oil tends to change
toward thermodynamic equilibrium during storage.
2.1. Composition and physicochemical properties
Bio-oils are multi-component mixtures of dier-
ent size molecules derived from depolymerization and
0196-8904/$ - see front matter 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2006.05.010
*
Corresponding author. Tel.: +86 20 87057760; fax: +86 20 87057789.
E-mail addresses: zhangqi@ms.giec.ac.cn (Q. Zhang), changjie@ms.
giec.ac.cn (J. Chang).
www.elsevier.com/locate/enconman
Energy Conversion and Management 48 (2007) 8792
fragmentation of cellulose, hemicellulose and lignin. There-
fore, the elemental composition of bio-oil and petroleum
derived fuel is dierent, and the basic data are shown in
Table 1.
2.1.1. Water
Bio-oil has a content of water as high as 1530 wt%
derived from the original moisture in the feedstock and
the product of dehydration during the pyrolysis reaction
and storage. The presence of water lowers the heating value
and ame temperature, but on the other hand, water
reduces the viscosity and enhances the uidity, which is
good for the atomization and combustion of bio-oil in
the engine. Shihadeh and Hochgreb [7] compared the bio-
oils of NREL (National Renewable Energy Laboratory,
US) to those of ENSYN (Ensyn Technologies, Inc., CA)
and found that additional thermal cracking improved its
chemical and vaporization characteristics. The better per-
formance and better ignition of NREL oil derived from
its lower water content and lower molecular weight.
2.1.2. Oxygen
The oxygen content of bio-oils is usually 3540%[6,8], dis-
tributed in more than 300 compounds depending on the
resource of biomass and severity of the pyrolytic processes
(temperature, residence time and heating rate). The presence
of oxygen creates the primary issue for the dierences
between bio-oils and hydrocarbon fuels. The high oxygen
content leads to the lower energy density than the conven-
tional fuel by 50% and immiscibility with hydrocarbon fuels
also. In addition, the strong acidity of bio-oils makes them
extremely unstable. Because of their complex compositions,
bio-oils show a wide range of boiling point temperature.
During the distillation, the slowheating induces the polymer-
ization of some reactive components, and bio-oils start boil-
ing below 100 C, while stopping at 250280 C, leaving 35
50 wt% as solid residues. Therefore, bio-oils cannot be used
in the instance of complete evaporation before combustion.
2.1.3. Viscosity
Depending on the biomass feedstocks and pyrolytic
processes, the viscosities of bio-oils vary in a large range.
Bio-oils [9] produced from Pterocarpus indicus and Fraxi-
nus mandshurica had a kinetic viscosity of 70350 mPa s
and 1070 mPa s, respectively, and that from rice straw
had a minimum kinetic viscosity of about 510 mPa s for
its high water content.
Sipilae` et al. [10] investigated the bio-oils from hard-
wood, softwood and straw by ash pyrolysis in an atmo-
spheric uidized bed. It was found that the viscosities
were reduced in the bio-oils with higher water content
and less water insoluble components. Viscosity was also
aected by alcohols: an addition of 5 wt% methanol into
hardwood pyrolysis oil with low methanol content
decreased its viscosity by 35%. The straw oil is less viscous
and had the highest methanol content of 4 wt%.
The research [11] of NREL showed that the viscosity
increased only from 20 to 22 cP over a 4 month period
when stored at 20 C with 10% methanol addition to the
bio-oil. This would extrapolate to a viscosity of 30 cP after
storage for 12 months. Ethanol at 20% had a similar stabi-
lizing eect. With 10viscosity at 40 C rose from about
13 cP to an interpolated 15 cP after preheating for 12 h at
90 C, e.g. to reduce the viscosity for ease of atomization.
Boucher et al. [12] tested bio-oil performance with the
addition of methanol regarding its use as a fuel for gas tur-
bine applications. The methanol reduced the density and
viscosity and increased the stability with the limitation of
a lowered ash point in the blend.
2.1.4. Acidity
Bio-oils comprise substantial amounts of carboxylic
acids, such as acetic and formic acids, which leads to low
pH values of 23. The bio-oil of pine had a pH of 2.6, while
that of hardwood was 2.8 [10]. Acidity makes bio-oil very
corrosive and extremely severe at elevated temperature,
which imposes more requirements on construction materi-
als of the vessels and the upgrading process before using
bio-oil in transport fuels.
2.1.5. Heating value
The properties of bio-oils depend on factors, such as
biomass feedstocks, production processes, reaction condi-
tions and collecting eciency. Usually the bio-oils of oil
plants have a higher heating value compared with those
of straw, wood or agricultural residues. Beis et al. [13] con-
ducted pyrolysis experiments on a sample of saower seed
and obtained bio-oil with a heating value of 41.0 MJ/kg
and a maximum yield of 44%. Ozcimen and Karaosmano-
glu [14] produced bio-oil from rapeseed cake in a xed bed
with a heating value of 36.4 MJ/kg and a yield of 59.7%.
However, taking wood and agricultural residues as raw
materials, the bio-oils have a heating value of about
20 MJ/kg and a yield up to 7080%.
2.1.6. Ash
The presence of ash in bio-oil can cause erosion, corro-
sion and kicking problems in the engines and the valves
and even deterioration when the ash content is higher than
Table 1
Typical properties of wood pyrolysis bio-oil and of heavy fuel oil [6]
Physical property Bio-oil Heavy fuel oil
Moisture content (wt%) 1530 0.1
pH 2.5
Specic gravity 1.2 0.94
Elemental composition (wt%)
C 5458 85
H 5.57.0 11
O 3540 1.0
N 00.2 0.3
Ash 00.2 0.1
HHV (MJ/kg) 1619 40
Viscosity (at 50 C) (cP) 40100 180
Solids (wt%) 0.21 1
Distillation residue (wt%) up to 50 1
88 Q. Zhang et al. / Energy Conversion and Management 48 (2007) 8792
0.1 wt%. However, alkali metals are problematic compo-
nents of the ash. More specically, sodium, potassium
and vanadium are responsible for high temperature corro-
sion and deposition, while calcium is responsible for hard
deposits. The H50 bio-oil was found to contain 2 ppm K,
6 ppm Na and 13 ppm Ca [12]. The best job of hot gas l-
tering to date at NREL [15] resulted in less than 2 ppm
alkali metals and 2 ppm alkaline earth metals in the bio-oil.
2.2. Compositions of bio-oil
The 99.7% of bio-oil, a complex mixture containing car-
bon, hydrogen and oxygen, is composed of acids, alcohols,
aldehydes, esters, ketones, sugars, phenols, guaiacols,
syringols, furans, lignin derived phenols and extractible ter-
pene with multi-functional groups [16].
Wang et al. [17] analyzed the compositions of bio-oils
from F. mandshurica pyrolysis by gas chromatography-
mass spectroscopy (GCMS) and illustrated the similarities
of the main contents, that is that the fragments, such as fur-
fural, dimethyl phenol, 2-methoxy-4-methyl phenol, euge-
nol, cedrol, furanone, etc., are a large proportion in every
bio-oil. Most of the components identied are the phenols
with ketones and aldehydes groups attached, and nearly all
the functional groups showed the extensive existence of
oxygen. On the other hand, the analysis proved that the
abundant aldehydes and ketones make bio-oils especially
hydrophilic and highly hydrated, which leads to the water
being dicult to eliminate.
The bio-oil from P. indicus [9] is mainly comprised of
levoglucosan, furfural, phenols (with methyl, methoxy
and/or propenyl groups), aldehydes (including benzalde-
hyde with methyl and/or hydroxyl) and vanillin according
to the GCMS analysis. Table 2 shows the distribution of
some detected compounds in the organic composition of
bio-oil from P. indicus.
The main components of the bio-oil obtained from
spruce wood in Adams [18] study exhibited many similar
components with those in Table 2.
Zhang et al. [19] separated the bio-oil into four frac-
tions: aliphatic, aromatic, polar and non-volatile fragments
by using solvent extraction and liquid chromatography on
an aluminum column. Identication revealed the high con-
tents of acetic acids and hydroxyacetones in the water
phase and more polar components and less aliphatic and
aromatic hydrocarbons in the oil phase.
In conclusion, bio-oils are a complex mixture, highly-
oxygenated with a great amount of large size molecules,
which nearly involve all species of oxygenated organics,
such as esters, ethers, aldehydes, ketones, phenols, carbox-
ylic acids and alcohols [20].
3. Upgrading of bio-oil
The deleterious properties of high viscosity, thermal
instability and corrosiveness present many obstacles to
the substitution of fossil derived fuels by bio-oils. So, an
upgrading process by reducing the oxygen content is
required before its application. The recent upgrading tech-
niques are described as follows.
3.1. Hydrodeoxygenation
The hydro-process is performed in hydrogen providing
solvents activated by the catalysts of CoMo, NiMo
and their oxides or loaded on Al
2
O
3
under pressurized con-
ditions of hydrogen and/or CO. Oxygen is removed as H
2
O
and CO
2
, and then the energy density is elevated. Pindoria
[21,22] hydrotreated the volatiles from fast pyrolysis of
eucalyptus in a two stage reactor. Hydro-cracking without
catalysts was operated in the rst stage, and catalytic
hydrotreatment was operative in the second stage with
lower temperature and the same pressure compared with
that in the rst stage. The analysis indicated that the deac-
tivation of the catalyst did not result from carbon deposi-
tion. Instead, the embodiments of volatile components
blocked the activated sites in the zeolite catalyst. This
hydro-process produced much water and complicated the
bio-oil with many impurities.
Zhang et al. [23] separated the bio-oil with a yield of
70% into water and oil phases, and the oil phase was
hydrotreated and catalyzed by sulphided CoMoP/
Al
2
O
3
. The reaction was operated in an autoclave lled
with tetralin (as a hydrogen donor solvent) under the opti-
mum conditions of 360 C and 2 MPa of cold hydrogen
pressure. The oxygen content was reduced from 41.8% of
the crude oil to 3% of the upgraded one, and besides, the
Table 2
Relative content of main compounds in organic composition of bio-oil
produced from P. indicus
Compound Relative content (%)
Furfural 9.06
Acetoxyacetone, 1-hydroxyl 1.21
Furfural, 5-methyl 1.82
Phenol 2.55
2-Cyclopentane-1-one, 3-methyl 1.58
Benzaldehyde, 2-hydroxyl 2.70
Phenol, 2-methyl 5.04
Phenol, 4-methyl 0.51
Phenol, 2-methoxyl 0.27
Phenol, 2,4-dimethyl 9.62
Phenol, 4-ethyl 2.18
Phenol, 2-methoxy-5-methyl 4.15
Phenol, 2-methoxy-4-methyl 0.55
Benzene, 1,2,4-trimethoxyl 3.80
Phenol, 2,6-dimethyl-4-(1-propenyl) 4.25
1,2-Benzenedicarboxylic acid, diisooctyl ester 1.80
2-Furanone 5.70
Levoglucosan 6.75
Phenol, 2,6-dimethoxy-4-propenyl 3.14
Furanone, 5-methyl 0.49
Acetophenone, 1-(4-hydroxy-3-methoxy) 2.94
Vanillin 6.35
Benzaldehyde, 3,5-dimethyl-4-hydroxyl 4.54
Cinnamic aldehyde, 3,5-demethoxy-4-hydroxyl 2.19
Q. Zhang et al. / Energy Conversion and Management 48 (2007) 8792 89
crude oil was methanol soluble while the upgraded one was
oil soluble for the dehydroxy.
Elloit and Neuenschwander [24] examined the hydro-
catalytic reactions of bio-oils from NREL, UF (Union
Electrica Fenosa) and RPT3 (Ensyn Technologies, Inc.)
in a continuous feed xed bed reactor. Higher conversion
was obtained in the downow conguration compared to
the upow one. The conversion was doubled on NiMo/c-
Al
2
O
3
compared to the CoMo/spinel at elevated
temperature.
Senol et al. [25] eliminated oxygen in carboxylic groups
with model compounds of methyl heptanoate and methyl
hexanoate on sulphided NiMo/c-Al
2
O
3
and CoMo/c-
Al
2
O
3
in a ow reactor to discern the reaction schemes. Ali-
phatic methyl esters produced hydrocarbons via three main
pathways: the rst pathway gave alcohols followed by
dehydration to hydrocarbons. The de-esterication yielded
an alcohol and a carboxylic acid in the second pathway.
Carboxylic acid was further converted to hydrocarbons
either directly or with an alcohol intermediate.
Apparently the hydro-treating process needs compli-
cated equipments, superior techniques and excess cost
and usually is halted by catalyst deactivation and reactor
clogging.
3.2. Catalytic cracking of pyrolysis vapors
Oxygen containing bio-oils are catalytically decomposed
to hydrocarbons with the removal of oxygen as H
2
O, CO
2
or CO. Nokkosmaki et al. [26] proved ZnO to be a mild
catalyst on the composition and stability of bio-oils in
the conversion of pyrolysis vapors, and the liquid yields
were not found to be substantially reduced. Although it
had no eect on the water insoluble fraction (lignin
derived), it decomposed the diethyl insoluble fraction
(water soluble anhydrosugars and polysaccharides). After
heating at 80 C for 24 h, the increase in viscosity was sig-
nicantly lowered for the ZnO-treated oil (55% increase in
viscosity) compared to the reference oil without any cata-
lyst (129% increase in viscosity). Despite the indicated
deactivation of the catalyst, the improvement in the stabil-
ity of the ZnO treated oil was clearly observed.
Adam et al. [18] illustrated the eects and catalytic prop-
erties of Al-MCM-41, Cu/Al-MCM-41 and Al-MCM-41
with pores enlarged on bio-oil upgrading. The resulting
compositions of vapors were changed through the catalysts
layer. Levoglucosan was completely eliminated, which was
typical for each catalyst. While the catalysis increased the
yields of acetic acid, furfural and furans, it reduced those
of large molecular phenols among the cellulose pyrolysis
products of spruce wood. The pore size enlargement and
incorporation of catalysts reduced the yield of acetic acid
and water.
Adjaye and Bakhshi [27,28] studied the catalytic perfor-
mance of the dierent catalysts for upgrading of bio-oil.
Among the ve catalysts studied, HZSM-5 was the most
eective catalyst for producing the organic distillate
fraction, overall hydrocarbons and aromatic hydrocarbons
and the least coke formation. Reaction pathways were pos-
tulated that bio-oil conversion proceeded as a result of
thermal eects followed by thermocatalytic eects. The
thermal eects produced separation of the bio-oil into light
organics and heavy organics and polymerization of the bio-
oil to char. The thermocatalytic eects produced coke, tar,
gas, water and the desired organic distillate fraction.
Guo et al. [29] reviewed various catalysts used in bio-oil
upgrading in detail and believed that although catalytic
cracking is a predominant technique, the catalyst with
good performance of high conversion and little coking ten-
dency is demanding much eort.
Although catalytic cracking is regarded as a cheaper
route by converting oxygenated feedstocks to lighter frac-
tions, the results seem not promising due to high coking
(825 wt%) and poor quality of the fuels obtained.
3.3. Emulsication
The simplest way to use bio-oil as a transport fuel seems
to be to combine it with Diesel directly. Although the bio-
oils are immiscible with hydrocarbons, they can be emulsi-
ed by the aid of a surfactant. Chiaramonti [30,31] pre-
pared emulsied bio-oil by the ratios of 25, 50 and
75 wt% and found the emulsions were more stable than
the original ones. The higher the bio-oil content, the higher
the viscosity of the emulsion. The optimal range of emulsi-
er to provide acceptable viscosity was between 0.5% and
2%. In particular, the eect of the long term use of emul-
sions on the stainless steel engine and its subassemblies
should be estimated.
Ikura et al. [32] obtained light fractions of bio-oil by
centrifugation and emulsied them in No. 2 Diesel by
CANMET surfactant with ratios of 10, 20 and 30 wt% sep-
arately. The cost of producing stable emulsions were
2.6 cents/L for 10% emulsion, 3.4 cents/L for 20% emul-
sion and 4.1 cents/L for 30% emulsion separately. The
bio-oil was determined to have a cetane number of 5.6,
which will decrease by 0.4 for each 10% concentration aug-
mentation. The viscosity of 1020 wt% emulsions was
much lower than that of pure bio-oil, and their corrosive-
ness was about half that of bio-oil alone.
Emulsication does not demand redundant chemical
transformations, but the high cost and energy consumption
input cannot be neglected. The accompanying corrosive-
ness to the engine and the subassemblies is inevitably
serious.
3.4. Steam reforming
Hydrogen is a clean energy resource and very important
in the chemical industry, and the rising focus on reforming
the water fraction of bio-oil looks promising. Production
of hydrogen from reforming bio-oil was investigated by
NREL [33,34] extensively, including the reactions in a xed
bed or a uidized bed and studies of the reforming
90 Q. Zhang et al. / Energy Conversion and Management 48 (2007) 8792
mechanisms on model compounds. The xed bed used in
the conventional reforming of natural gas or naphtha is
not suitable in the lignin derived fraction of bio-oil because
of its tendency to decompose thermally and form carbon
deposits on the upper layer of the catalyst and in the reac-
tor freeboard.
Czernik et al. [35] obtained hydrogen in a uidized bed
reactor from the carbohydrate derived fraction of wood
pyrolysis oil with a yield of about 80% of theoretical, which
corresponds to approximately 6 kg of hydrogen from100 kg
of wood used for pyrolysis. Commercial nickel catalysts
showed good activity in processing biomass derived liquids
and was readily regenerated (20 min to 2 h) by steam or
CO
2
gasication after deactivation, which occurred during
reforming. The commercial catalysts, designed for xed
bed applications, were susceptible to attrition in the uidized
bed. Consequently, they were entrained at a rate of 5%/day.
The development of a uidizable catalyst that has both high
activity and mechanical strength at the conditions of the
steam reforming process is needed and is being pursued.
Garcia et al. [36] chose magnesium and lanthanum as
support modiers to enhance steam adsorption that facili-
tates the carbon gasication, while cobalt and chromium
additives were applied to alleviate the coke formation reac-
tions, which modied the metal sites forming alloys with
nickel and possibly reducing the crystallite size. More
hydrophilic sites surrounding nickel crystallites were eec-
tive in extending the duration of the catalysts activity. The
catalyst deactivation upon steam reforming of complex
bio-oils occurs much faster than with natural gas or naph-
tha, but the catalyst can be eciently regenerated by steam
or CO
2
gasication.
Takanabe et al. [37] completely converted acetic acid,
the model compound, by steam reforming over Pt/ZrO
2
catalysts and found a hydrogen yield close to thermody-
namic equilibrium. Analysis showed that Pt was essential
for the steam reforming to proceed, and ZrO
2
was needed
to activate the steam, which was also active for oligomer
precursor formation under the conditions investigated.
The results illustrated that steam reforming took place at
the PtZrO
2
boundary, and that deactivation occurred
when this boundary is blocked by oligomers.
3.5. Chemicals extracted from the bio-oils
Hundreds of the components of bio-oil are determined,
and reclaiming one or more kinds of small and valuable
chemicals arouses great interests among scholars and busi-
nessmen. There are many substances that can be extracted
from bio-oil, such as phenols used in the resins industry,
volatile organic acids in formation of de-icers, levogluco-
san, hydroxyacetaldehyde and some additives applied in
the pharmaceutical, ber synthesizing or fertilizing indus-
try and avoring agents in food products [1,38]. Commer-
cialization of special chemicals from bio-oils requires more
devotion to developing reliable low cost separation and
rening techniques.
4. Conclusions and recommendations for future work
Bio-oils have received extensive recognition from inter-
national energy organizations around the world for their
characteristics as fuels used in combustors, engines or gas
turbines and resources in chemical industries. Some prob-
lems blocking its industrialization processes and recom-
mendations are described as follows:
Since bio-oils are complex and chemically unstable mix-
tures, biomass fast pyrolysis should corresponded with
the characteristics of the biomass feedstocks, products
uses, suitable reactors and processes for the products
applications.
Much more work is needed on the stabilization and
upgrading of bio-oils with some modications to equip-
ment conguration before applying them in generating
heat or power. It is important to learn more about the
mechanisms involved in adding solvent to slow the
apparent aging period.
Hydrodeoxygenation, catalytic cracking and steam
reforming of bio-oils are so complicated techniques that
steady, dependable, fully developed reactors are
urgently desirable.
According to the main components and reasonable reac-
tions, it is meaningful to envision adding some solvents
to reduce polymerization by esterication, acetalization
and phenol/formaldehyde reactions, antioxidants to
reduce olen polymerization reactions and emulsiers
to prevent phase separation problems.
Biomass can be an economically viable renewable
resource for ecient energy utilities if used in an inte-
grated process that also generates other marketable
coproducts to maintain sustainable developments.
Acknowledgements
The supports from National natural science foundation
of China (50476090) and Guangdong natural science foun-
dation (04000378) are greatly appreciated.
References
[1] Bridgwater AV, Peacocke GVC. Fast pyrolysis processes for biomass.
Sustain Renew Energy Rev 2000;4:173.
[2] Scott DS, Piskorz J, Radlein D. Liquid products from the continuous
ash pyrolysis of biomass. Ind Eng Chem Process Des Dev 1985;
24:5816.
[3] Czernik SR, Bridgwater AV. Overview of applications of biomass fast
pyrolysis oil. Energy Fuels 2004;18(2):5908.
[4] Moore A. Environmentally sound production of liquid biofuels: the
view of the European Commission. Proceedings of the International
workshop on environmental aspects of energy crop production,
Brasimone, Italy, October 910, 1997. Wieselburg: BLT; 1998. p.
4367.
[5] Demirbas A. Biomass resource facilities and biomass conversion
processing for fuels and chemicals. Energy Convers Manage 2001;42:
135778.
Q. Zhang et al. / Energy Conversion and Management 48 (2007) 8792 91
[6] Oasmaa A, Czernik S. Fuel oil quality of biomass pyrolysis oils-state
of the art for the end-users. Energy Fuels 1999;13:91421.
[7] Shihadeh A, Hochgreb S. Impact of biomass pyrolysis oil process
conditions on ignition delay in compression ignition engines. Energy
Fuels 2002;16:55261.
[8] Scholze B, Meier D. Characterization of the water-insoluble fraction
from pyrolysis oil (pyrolytic lignin) Part I. PY-GC/MS, FTIR, and
functional groups. J Anal Appl Pyrol 2001;60:4154.
[9] Luo ZY, Wang SR, Liao YF, et al. Research on biomass fast
pyrolysis for liquid fuel. Biomass Bioenergy 2004;26:45562.
[10] Sipilae` K, Kuoppala E, Fagernae`s L, et al. Characterization of
biomassbased ash pyrolysis oils. Biomass Bioenergy 1998;14(2):
10313.
[11] Diebold JP. A review of the chemical and physical mechanisms of the
storage stability of fast pyrolysis bio-oil. Available from: http://
webdev.its.iastate.edu/webnews/data/site_biorenew_reading/19/web-
newsleeld_le/ReviewOfMechanisms.pdf [2005-08-20].
[12] Boucher ME, Chaala A, Roy C. Bio-oils obtained by vacuum
pyrolysis of softwood bark as a liquid fuel for gas turbines. Part I:
Properties of bio-oil and its blends with methanol and a pyrolytic
aqueous phase. Biomass Bioenergy 2000;19:33750.
[13] Beis SH, Onay O, Kockar OM. Fixed-bed pyrolysis of saower seed:
inuence of pyrolysis parameters on product yields and compositions.
Renew Energy 2002;26:2132.
[14] Ozcimen D, Karaosmanoglu F. Production and characterization of
bio-oil and biochar from rapeseed cake. Renew Energy 2004;29:
77987.
[15] Scahill J, Diebold JP, Feik C. Removal of residual char nes from
pyrolysis vapors by hot gasication. In: Bridgewater AV, Bock DGB,
editors. Developments in thermochemical biomass conversion. Lon-
don: Blackie Academic and Professional; 1996. p. 25366.
[16] Guo Y, Wang Y, Wei F, et al. Research progress in biomass ash
pyrolysis technology for liquids production. Chem Ind Eng Progr
2001;8:137.
[17] Wang SR, Luo ZY, Tan H, et al. The analyses of characteristics bio-
oil produced from biomass ash pyrolysis. J Eng Thermophys 2004;
25(6):104952.
[18] Adam J, Blazso M, Meszaros E, et al. Pyrolysis of biomass in the
presence of Al-MCM-41 type catalysts. Fuel 2005;84(1213):
1494502.
[19] Zhang SP, Yan YJ, Ren ZW, et al. Analysis of liquid product
obtained by the fast pyrolysis of biomass. J Chin Sci Technol 2001;
27:6668.
[20] Peng WM, Wu QY. Production of fuels from biomass by pyrolysis.
New Energy Sour 2000;22(11):3944.
[21] Pindoria RV, Lim J-Y, Hawkes JE, et al. Structural characterization
of biomass pyrolysis tars/oils from eucalyptus wood wastes: eect of
H
2
pressure and samples conguration. Fuel 1997;76(11):101323.
[22] Pindoria RV, Megaritis A, Herod AA, et al. A two-stage xed-bed
reactor for direct hydrotreatment of volatiles from the hydropyrolysis
of biomass: eect of catalyst temperature, pressure and catalyst
ageing time on product characteristics. Fuel 1998;77(15):171526.
[23] Zhang SP, Yan Yongjie, Li T, et al. Upgrading of liquid fuel from
the pyrolysis of biomass. Bioresour Technol 2005;96:54550.
[24] Elliott DC, Neuenschwander GG. Liquid fuel by low-severity
hydrotreating of biocrude. Developments in thermochemical biomass
conversion. London: Blackie Academic and Professional; 1996. p.
61121.
[25] Senol OI, Viljava TR, Krause AOI. Hydrodeoxygenation of methyl
esters on sulphided NiMo/c-Al
2
O
3
and CoMo/c-Al
2
O
3
catalysts.
Catal Today 2005;100(34):3315.
[26] Nokkosmaki MI, Kuoppala ET, Leppamaki EA, et al. Catalytic
conversion of biomass pyrolysis vapours with zinc oxide. J Anal Appl
Pyrol 2000;55:11931.
[27] Adjaye JD, Bakhshi NN. Production of hydrocarbons by catalytic
upgrading of a fast pyrolysis bio-oil. part I: conversion over various
catalysts. Fuel Process Technol 1995;45(3):16183.
[28] Adjaye JD, Bakhshi NN. Production of hydrocarbons by catalytic
upgrading of a fast pyrolysis bio-oil. Part II: Comparative catalyst
performance and reaction pathways. Fuel Process Technol 1995;
45(3):185202.
[29] Guo XY, Yan YJ, Ren ZW. The using and forecast of catalyst in bio-
oil upgrading. Acta Energiae Solaris Sin 2003;124(12):20612.
[30] Chiaramonti D, Bonini M, Fratini E, et al. Development of
emulsions from biomass pyrolysis liquid and Diesel and their use in
enginesPart 1: Emulsion production. Biomass Bioenergy 2003;25:
8599.
[31] Chiaramonti D, Bonini M, Fratini E, et al. Development of
emulsions from biomass pyrolysis liquid and Diesel and their use in
enginesPart 2: Tests in Diesel engines. Biomass Bioenergy 2003;25:
10111.
[32] Ikura M, Stanciulescu M, Hogan E. Emulsication of pyrolysis
derived bio-oil in Diesel fuel. Biomass Bioenergy 2003;24:22132.
[33] Wang D, Czernik S, Montane D, et al. Biomass to hydrogen via
pyrolysis and catalytic steam reforming of the pyrolysis oil and its
fractions. Ind Eng Chem Res 1997;36:150718.
[34] Wang D, Czermik S, Chornet E. Production of hydrogen from
biomass by catalytic steam reforming of fast pyrolytic oils. Energy
Fuels 1998;12:1924.
[35] Czernik S, French R, Feik C, et al. Hydrogen by catalytic steam
reforming of liquid byproducts from biomass thermoconversion
processes. Ind Eng Chem Res 2002;41:420915.
[36] Garcia L, French R, Czernik S, et al. Catalytic steam reforming of
bio-oils for the production of hydrogen: eects of catalyst composi-
tion. Appl Catal A: Gen 2000;201:22539.
[37] Takanabe K, Aika K, Seshan K, et al. Sustainable hydrogen from
bio-oilsteam reforming of acetic acid as a model oxygenate. J Catal
2004;227:1018.
[38] Bridgwater AV, Meier D, Radlein D. An overview of fast pyrolysis of
biomass. Org Geochem 1999;30:147993.
92 Q. Zhang et al. / Energy Conversion and Management 48 (2007) 8792

Das könnte Ihnen auch gefallen