Sie sind auf Seite 1von 13

Review

pubs.acs.org/IECR

Bio-oils Upgrading for Second Generation Biofuels


Inês Graça,† José M. Lopes,†,* Henrique S. Cerqueira,‡ and Maria F. Ribeiro†

IBB, Institute for Biotechnology and Bioengineering, Centre for Biological and Chemical Engineering, Instituto Superior
Técnico/Technical University of Lisbon, Av. Rovisco Pais, 1049-001 Lisboa, Portugal

OSX, Praça Mahatma Gandhi, 14/11th floor, 20031-100 Rio de Janeiro, RJ, Brazil

ABSTRACT: The envisaged upgrading of lignocellulosic biomass derived feedstocks (bio-oils) in dedicated units or by
coprocessing in existing units of the refinery, to partially replace crude oil in the production of transportation fuels, is a topic that
has been receiving much attention from both industry and academia in recent years. Regardless of lignocellulosic biomass origin,
these feedstocks are complex mixtures of many oxygenated hydrocarbons. Therefore, their upgrading toward liquid fuels must
include oxygen removal. So far, two main routes have been proposed, considering many studies at laboratory scale and others
from industry: catalytic hydrotreatment (HDT), mainly hydrodeoxygenation (HDO), and catalytic cracking, technologies that
are already present in today’s refineries configuration. HDO has been performed at high hydrogen pressure, using catalysts based
on those typically applied in conventional hydrotreating, as well as a new type of supported noble metal and transition metal
catalysts. Catalytic cracking occurs at atmospheric pressure, using acid catalysts, mainly the active phases of fluid catalytic cracking
(FCC) catalysts (HY and HZSM-5 zeolites). The present review is then focused on the upgrading possibilities of renewable
nonedible feedstocks, obtained from biomass fast pyrolysis or liquefaction, in petroleum refineries, toward the production of
second generation biofuels. It includes the recent studies concerning the alternative of bio-oils coprocessing together with crude
oil feedstocks. In fact, although all these feedstocks have the potential to be directly converted into transportation fuels in
dedicated units, it seems more attractive to upgrade them in combination with conventional feedstocks.

1. INTRODUCTION biopolymers, cellulose (glucose polymer), hemicellulose (shorter


The partial replacement of crude oil by renewable sources to polymers of various sugars, principally xylose), and lignin
produce high quality transportation fuels is a timely and in- (propyl-phenol polymer).3,8,9 These three essential compo-
teresting topic. It helps to reduce the dependence from petroleum nents form an insoluble three-dimensional network with a very
as well as carbon dioxide emissions responsible for the greenhouse complex structure, which makes the lignocellulosic biomass
effect, due to its CO2 neutrality.1,2 Contrarily to heavy oil, biomass more difficult to convert than sugars or vegetable oils.3
presents the advantage of containing negligible amounts of sulfur, Pyrolysis and liquefaction are the two main processes devel-
nitrogen, and metals and is easily available in many countries. oped to directly convert biomass feedstocks into liquid products,
Moreover, although several renewable sources, such as biomass, generally called bio-oils.10−14 Whatever the process, this trans-
wind, solar, and hydropower, can be used to generate heat and formation is carried out in one single reactor and a very large
power, biomass is currently the only renewable source of carbon fraction of the biomass energy (50−90%) can be converted into
that can be converted into liquid fuels,3−5 which can be used in the liquids. Fast pyrolysis technology15 involves short residence
actual infrastructure of the transportation sector.1 times (typically less than 2 s), fast heating rates, high tem-
Many types of biomass feedstocks are available for the peratures, in the range of 400−700 °C, and low pressures (1−5
production of biofuels, namely sugars, vegetable oils (“first atm), in an oxygen-free atmosphere. On the other hand, liquefac-
generation” biofuels) and lignocellulosic biomass, as residues tion occurs at high hydrogen pressure (50−200 bar), in a lower
from agriculture and forestry (“second generation” biofuels).3 temperature range of 250−325 °C. Thus, pyrolysis presents a
In recent years, first generation biofuels (bioethanol and bio- lower capital cost than liquefaction, favoring its commercialization.1
diesel) have been increasing their market share, mainly because These biomass derived oils constitute an alternative energy
of their blending with conventional fuels. However, it should be source, but they cannot be directly used as transportation fuels.
mentioned that lignocellulosic biomass can be grown in com- They are complex mixtures of more than 400 different oxygenated
bination with food or on nonagricultural lands, therefore not hydrocarbons, as carboxylic acids, aldehydes, alcohols, ketones,
compromising the sustainable utilization of fertile land, that is, esters, ethers, phenols, furans and carbohydrates. Mainly due to
not competing with food production.1,6 Lignocellulosic biomass their oxygen-rich composition, bio-oils have some undesired prop-
consists of an abundant renewable source of carbon,7 con- erties for fuel applications: low heating value (less than 50% of that
vertible into liquid carburant, which could be introduced in an of conventional fuel oils), immiscibility with hydrocarbon fuels,
existing downstream process, increasing the production of com- thermal and chemical instability, high density, corrosiveness, etc.1,9,16,17
mercial fuels in a sustainable manner.
The key difference between petroleum and renewable raw Received: June 28, 2012
materials is the higher oxygen content of the latter: residues Revised: November 14, 2012
from biomass typically contain between 35 and 50 wt % oxygen.1 Accepted: November 20, 2012
Lignocellulosic materials comprise three different types of Published: November 21, 2012

© 2012 American Chemical Society 275 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

The bio-oils produced are usually dark brown organic liquids hydrotreating and there is a wider range of possibilities to be
that have a distinctive smoky odor.18,19 Typically, the fast explored with mono- and multifunctional catalysts.11
pyrolysis oils present higher oxygen content, higher acidity, and Both upgrading technologies present important operating
lower heating value than liquefaction oils. Independently of the problems, such as reactor plugging, rapid catalyst deactivation,
origin, the bio-oils must undergo upgrading treatments in order and low liquid product yields, due to the bio-oil tendency to poly-
to increase their stability and to become commercially useful. merize even under mild heating. To overcome or minimize these
They require chemical transformations to increase volatility and problems, the possibility of coprocessing the bio-oils with con-
thermal stability and reduce viscosity through oxygen removal ventional feedstocks in existing refinery units was envisioned.34−36
and molecular weight reduction.20 In these cases, a previous partial hydrotreatment of the bio-oil
In 2007, the review by Huber and Corma21 reported the would be necessary in order to stabilize the feedstock and increase
increasing interest in biofuels production from biomass-derived its miscibility with petroleum fractions. In fact, because of the
feedstocks in petroleum refineries, thus taking advantage of above-mentioned chemical instability, small miscibility with hy-
economy of scale and experience by using existing infrastructures. drocarbons,35 and corrosiveness,37 its direct addition to estab-
It is also important to remark that many research efforts on the lished processes like FCC21,38 or HDT34 is not straightforward,
study of these possibilities have been, as well, carried out by several thus creating the need for at least one previous HDO upgrading
oil companies. step.
Two main routes for partial or total elimination of oxygen Therefore, the management of hydrogen plays a key role in
atoms have been proposed: catalytic cracking and catalytic eliminating the dependence upon external fossil fuels.39 It will
hydrotreating.21−26 The stoichiometric equations representative be desirable to minimize the hydrogen requirements, trying to
of these processes can be summarized as follows:27 couple strategies that consume less hydrogen with those that
Decarboxylation: produce hydrogen from renewable energy sources, such as solar,
wind or biomass.21
C6H8O4 → C4 H8 + 2CO2 Very recent reviews, from 2011 and 2012, reported the huge
research activity concerning these two refining bio-oils upgrad-
Hydrodeoxygenation: ing methods, HDO and catalytic cracking, in the perspective of
C6H8O4 + 4H 2 → C6H8 + 4H 2O dedicated units development.40−42
The aim of the present paper is to review the upgrading
Cracking: possibilities of renewable nonedible feedstocks, as lignocellulosic
biomass derived oils, by coprocessing or in dedicated refining
C6H8O4 → C4.5H8 + H 2O + 1.5CO2
units, toward the production of end-usable liquid fuels (second
The catalytic cracking of bio-oils removes oxygen by simu- generation biofuels).
ltaneous dehydration−decarboxylation reactions, using cracking
catalysts (zeolites, silica−alumina) at atmospheric pressure.28 2. HYDRODEOXYGENATION (HDO)
The oxygen can be eliminated as water, carbon monoxide, carbon Bio-oils hydrotreating is based on a modification, or extension,
dioxide, and short-chain acids. The oxygenated feedstock is con- of conventional HDT process existing in refineries. It differs
verted into a lighter hydrocarbons fraction, falling particularly from processing petroleum or coal liquids because of the im-
in the gasoline boiling range and containing high aromatics portance of HDO, as opposed to nitrogen (HDN) or sulfur
contents.21,28 removal (HDS), and naturally the term deoxygenation is assumed
The other method proposed is the catalytic hydrotreatment for upgrading biomass derived oils. In this process, oxygen is
(HDT) or hydrodeoxygenation (HDO), performed at high removed through a family of deoxygenation reactions, which
temperature, under high hydrogen pressure, in the presence of can be represented as follows:11
a heterogeneous catalyst.29,30 The reactions occurring are elim- C6H8O4 + 6H 2 → 6CH 2 + 4H 2O
ination of oxygen as water, elimination of nitrogen as ammonia,
and hydrogenation-hydrocracking of large molecules, leading to It should be mentioned that most of the HDO previous reports
the production of hydrocarbons in the diesel range. The reaction focused on model molecules representative of bio-oil fragment
conditions and the catalysts first tested (e.g., NiMo, CoMo sulfide- studies, while the majority of recent works focus on catalysts
based catalysts) were similar to those used in the petroleum refining development, the work being extended from small batch reactor
HDT processes. The degree of deoxygenation can be modulated tests, in universities, to demonstration scale processing, in
from simple stabilization, that is, elimination of more reactive petroleum refining laboratories.
functions such as carbonyl, olefins, and carboxyls, to complete Since the first Furimsky review on HDO,43 published in
refining with a maximal hydrocarbon yield that will be a function 1983, the growing concern about the upgrading of coal and
of the extent and how the oxygen atoms are removed.11 The biomass-derived liquids led to an expansion of HDO under-
residual amount of oxygenates is a function of the HDO standing studies. Maggi and Delmon24 published a review in
severity.31−33 1997, and many studies about the catalytic chemistry, kinetics,
Catalytic cracking has been considered as a cheaper alterna- and mechanisms of HDO reactions, using various model oxy-
tive.28,34 It offers significant processing and economic advantages genated compounds, were collected and reported in another
over catalytic hydrotreating because hydrogen is not required review by Furimsky in 2000.22 Classical HDT catalysts (sulfided
and operates under atmospheric pressure, thus reducing operat- CoMo and NiMo) have been the first and most extensively
ing costs. However, the poor hydrocarbons yields, the extensive tested ones. Their deactivation problems were identified and
coking observed during catalytic cracking of bio-oils, and the new and promising catalytic systems, based on noble metals
quality of the fuels obtained (significant amounts of phenolic supported on carbon, were announced.
compounds) are drawbacks that need to be overcome. Cracking The extensive work that has been undertaken over the past
of bio-oils with zeolite catalysts is less well developed than 25−30 years in the field of catalytic hydrotreating of biomass-derived
276 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

Table 1. Tested Catalysts, Reactivity, And Main Products Obtained in the HDO Transformations of Model Oxygenate
Compounds
oxygenate catalysts relative reactivity products
guaiacol31,32,44−46 CoMo/Al2O3 low phenol; benzene
Rh,Pd,Pt/ZrO2
Ni2P,Fe2P,MoP,WP/silica
anisole48,49 CoMo/Al2O3 high phenol; o-cresol; benzene
Ni−Cu/supports
phenol50−58 CoMo/Al2O3 most refractory cyclohexanol; cyclohexanona; cyclohexane; benzene
Ni/Y zeolite or silica
cresols50,59−61 CoMo/Al2O3 high, effective at mild conditions toluene; cyclohexane
Pt/carbon
aldehydes and ketones63−65 noble metals/zeolites high diphenylmethane; toluene
esters66,67 NiMo,CoMo/γ-Al2O3 high C3−C5 alkanes and alkenes; methanol; carboxylic acids
acids68,69 NiMo,CoMo/Al2O3 high alkanes
Pd/carbon
furans70−76,79 NiMo,CoMo/Al2O3 or carbon high C3−C4 alkanes and alkenes; cyclohexane
Pt/SiO2 or Al2O3 or TiO2

liquids is documented in papers from Elliot et al.,20 Mortensen it was noticed that, over supported nickel (1.5−20.3 wt %) on
et al.,40 Choudhary and Philips,41 and Bridgwater.42 either Y zeolite or silica catalysts, both hydrogenation and
The chemistry of HDO (and catalytic cracking) of bio-oils, hydrogenolysis reactions of phenol took place.50 Besides cyclo-
including reactions mechanisms and their relations to active hexanone and cyclohexanol, which resulted, respectively, from
sites in the catalysts, kinetic models, and deactivation, were partial and full phenol hydrogenation, also cyclohexane was
extensively reviewed by Mortensen et al.40 produced.50 High nickel loadings (∼20 wt %) and temperatures
2.1. Model Compounds HDO. The lab-scale studies tended to favor hydrogenolysis reaction, leading to benzene.
involving different oxygenate model compounds have con- Cresols are transformed mainly into toluene and cyclohexane
tributed to the understanding of HDO mechanisms. Most of over sulfided CoMo/Al2O3.50,59,61 Both direct HDO and a
these studies were performed in fixed-bed reactors, using pure transformation via hydrogenated phenol are possible.22 Noble
hydrogen (20−50 bar), at temperatures in the range 250−450 °C metal Pt/carbon catalysts were also recently shown to be active
and focused on catalyst development. Table 1 summarizes the in the deoxygenating of phenolic hydroxyl groups under mild
main data concerning the HDO reactivity of most relevant conditions, in the presence of diethylamine.62
oxygenates, as well as the tested catalysts and products obtained. The most recent studies on aldehydes and ketones HDO re-
Guaiacol revealed to be more refractory to deoxygenation ported that benzophenone and several aldehydes can effectively
than carbonyl and carboxylic groups31,32 and to have a higher be deoxygenated over noble metal supported zeolite catalysts.63−65
coking tendency. The proposed basic reaction scheme involves Benzophenone can be transformed into diphenylmethane over
the guaiacol hydrogenation into catechol and then into phenol. Pd/zeolites.63 Two mechanisms have been proposed: (i) direct
For the new catalyst formulations, monometallic or bimetallic hydrogenolysis of the CO bond or (ii) hydrogenation to
catalysts prepared by the impregnation of Rh, Pd, and Pt on alcohol followed by hydrogenolytic splitting of the C−OH bond.
ZrO2 support,44 similar or better activities than those of the The most active catalyst was Pd/beta. The same series of Pd
classical CoMo/Al2O3 catalyst were obtained and less carbon catalysts was used in the hydrodeoxygenation of aldehydes.64
deposition was found.44 More recently, transition metal phos- Benzaldehyde can be converted into toluene. 2-Methyl-2-pentenal
phides (Ni2P, Fe2P, MoP, and WP supported on silica) also HDO was achieved over Pt, Pd, and Cu catalysts supported on
showed an ability to convert guaiacol into benzene and phenol,45 silica.65 Pt/SiO2 was found to be the most active, whereas Cu/
with higher stability than CoMo/Al2O3 catalyst, due to lower SiO2 the less active.
carbon deposition.46 In the case of the cohydrotreatment of bio- Esters can be totally hydrodeoxygenated on sulfided NiMo
oils with petroleum cuts, the presence of sulfur on the HDO feed and CoMo supported on γ-alumina, at 250−300 °C and 15 bar
was thought to have a positive effect on the conversion of of hydrogen,66,67 leading to the formation of C5−C7 alkanes
molecules such as guaiacol.47 and alkenes, methanol and carboxylic acids.
The anisole HDO, a molecule presenting some similarities The transformation of acid compounds was investigated to a
with the compounds resulting from the lignin depolymerization lower extent. Carbon-supported Pd catalysts were effective in
during fast pyrolysis of wood, can be performed over classical the stearic acid transformation into heptadecane,68 while sulfided
HDT catalysts.48 The catalyst acidity plays an important role on NiMo/Al2O3 was revealed to be more active than the CoMo/
anisole conversion into phenol, o-cresol, and benzene. Novel Al2O3 catalyst in the deoxygenation of heptanoic acid as well as
Ni−Cu catalysts have been prepared with various supports and of methyl heptanoate.69
showed to be effective in the same transformation.49 Furans are present at relatively low contents in the bio-oils.
Phenol has also been extensively used as a model compound Nevertheless, they have been tested in many HDO studies
for bio-oil HDO,50−57 since it is the main final product of other because of their presence in the conventional petroleum feeds,
phenols (alkoxy- and methoxy-phenols) transformation and it coal and shale oil. Furans have then to be taken into considera-
presents the most difficult C−O bond to cleave, being, con- tion for HDT coprocessing option. The furan ring hydro-
sequently, particularly resistant to HDO. While typical CoMo/ genation occurs efficiently at typical conditions of hydro-
Al2O3 catalysts mainly promoted the aromatic ring hydrogenation,58 processing, with the formation of tetrahydrofuran, which suffers
277 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

HDO much faster than furan. In fact, classical CoMo and NiMo content in organic acids, aldehydes, ketones and ethers, whereas
supported on Al2O3 or carbon catalysts were extensively used phenols, aromatics and alkanes amounts were increased.
for furans HDO.70−75 Important products are C3 to C4, both Several institutions have been active in catalyst research and
paraffins and olefins,70 and butadiene.72 The HDO of development for the pyrolysis oil upgrading. For instance, Shell
benzofuran, which revealed to be a good model molecule for has been evaluating the use of carbon-supported Ru, in order to
HDO of a real feedstock, can yield o-ethylphenol and 2,3- upgrade bio-oil from thermal pyrolysis to an extension suitable
dihydrobenzofuran.74 Dibenzofuran can be transformed into for application in FCC,36 as will be further discussed in section
single-ring hydrocarbons, mainly cyclohexane.75 Detailed mech- 3.2. The oxygen content of the oil phase decreased from 28.0 wt %
anisms for furan, tetrahydrofuran, benzofuran, and dibenzofuran (dry basis), at 230 °C, to 15.5 wt %, at 330 °C, operating at
transformations are described elsewhere.22,76−78 More recent hydrogen pressures in the range of 200−290 bar. Since a high
studies report the use of Pt-supported catalysts,76,79 butane being hydrogen consumption was observed during the heating up
preferentially formed with Pt/TiO2. Other studies on benzofuran period, it has been proposed that a first saturation of double
HDO80 pointed out that reduced Mo and NiMO catalysts bonds and formation of alcohols from aldehydes and ketones
are able to activate benzofuran, promoting, in a first step, its occur, followed by hydrodeoxygenation/dehydration, at higher
hydrogenation. temperatures. Similar observations were made with model com-
2.2. Bio-oil HDO Upgrading. HDT removal of sulfur and pounds (section 2.1).
nitrogen from hydrocarbons is a well-established refinery Honeywell/UOP93 has developed a two-step process for the
process, but oxygen removal on the scale needed to upgrade conversion of bio-oils into diesel, aviation, and naphtha boiling
bio-oils is at development stage. The optimal performance of point range fuels, or fuel blending components, by two-stage
current industrial HDT catalysts for petroleum feedstocks cannot deoxygenation of the pyrolysis oil and separation of the products.
be reproduced over bio-oils HDO, preventing an immediate and Taking into account the goals outlined by the United States
easy coprocessing in the existing HDT processes. for the production of renewable fuels by 2017 (35 billion
The conditions used for the bio-oils upgrading by HDO gallons per year),94 the work about bio-oils upgrading focused
depend on their origin. It is possible to achieve a high HDO on catalytic hydroprocessing has continued at PNNL (Pacific
conversion of the high pressure liquefaction oils in a single Northwest National Laboratory). Various conversion technol-
step.81 For example, an oxygen removal higher than 95% was ogies for generating gasoline and diesel fuels from bio-oils have
reached from wood liquefaction oil, at 300 °C, over a sulfided been evaluated, as it was shown in a recent report,94 which studies
CoMo/Al2O3 catalyst,82 being its deactivation level dependent a design case to treat 2000 ton/day of hybrid poplar wood chips to
on the possible previous desalting treatment.83 The same produce 76 million gallons/year of gasoline and diesel. This design
catalysts showed, in other examples, to be able to promote case study involves several processing steps: a fast pyrolysis to
efficiently the pyrolysis oils upgrading.84,85 obtain a highly oxygenated liquid product; hydrotreating of the fast
Nevertheless, in the case of pyrolysis oils, due to the high pyrolysis oil into a stable hydrocarbon oil, with less than 2%
tendency of several O-compounds to polymerize, forming coke- oxygen; hydrocracking of the heavy portion of the stable hydro-
like products, less stable liquids are produced. Therefore, it was carbon oil; distillation of the hydrotreated and hydrocracked oil
proposed a first stabilization step to remove an important part into gasoline and diesel fuel blend stocks; and, finally, a steam
of the oxygen content.9,10,31,86−88 After this preliminary treat- reforming of the process off-gas and supplemental natural gas, to
ment, the HDO of the stabilized bio-oils becomes similar to produce hydrogen for the hydrotreating and hydrocracking steps
phenols HDO. This concept of two stage process was introduced (Figure 1). The hydrotreating section comprises two catalytic
by researchers from Pacific Northwest National Laboratory reaction stages: the first reactor stabilizes the bio-oil by mild hy-
(PNNL), in 1989.20 In this improved process, the oil is stabilized drotreatment (240 °C; 172 bar; LHSV = 1 h−1) and the second
in a first reactor, operating at temperatures lower than 300 °C one operates at higher temperature and lower space velocity
and under high pressure of hydrogen; then, it is fed to a second (370 °C; 138 bar; LHSV = 0.14 h−1).
reactor, where a deeper deoxygenation is achieved at higher Despite promising, several studies about early two stage
temperatures (350−400 °C).89 Typical petroleum hydrotreating processes concluded that the process is too expensive because
catalysts (NiMo or CoMo supported on γ-alumina) were used. of the large amount of hydrogen consumed, the low product
Other examples of two stage treatments have been reported, yields, the need to further upgrade the products in a refinery
namely concerning the upgrading of bio-oils from olive oil and the corrosiveness of the raw oil.89 In fact, based on different
industry wastes,90 and from cellulose.91 In the case of cellulose scenarios for hydrogen-purchase fast pyrolysis and upgrading,
hydropyrolysis, single and two-stage treatments have been tested, techno-economic analyses are being performed,95 but the con-
in a fixed bed reactor. In the single stage tests, a FeS catalyst was clusions about high costs indicate that some aspects of the bio-
used; for the two-stage tests, the primary oils were passed over a oils upgrading processes require further developments. The
commercial NiMo/alumina catalyst, at 400 °C, where an costs are not acceptable for petroleum refiners, and some reports
additional reduction of the oxygen content and aromaticity of suggest that lower costs could be achieved by sending partially
the oils were registered, thus producing a lighter oil.91 deoxygenated materials to a refinery. It means that reducing the
Recent developments, regarding bio-oils upgrading through severity of hydrodeoxygenation, leaving about 7 wt % oxygen in
HDO, report also a noble-metal based catalysts promotion of the pyrolysis oil and avoiding aromatics hydrogenation, would
the hydrotreatment.92 Ru/C, Ru/TiO2, Ru/Al2O3, Pt/C, and reduce hydrogen consumption and hydrotreating capital costs.
Pd/C catalysts have been tested comparatively to the classical In Europe, a large research Project (BioCoup) has been devel-
sulfided NiMo/Al2O3 and CoMo/Al2O3, at 250 and 300 °C, oped that has as a major objective to arrive at pyrolysis oil-
and pressures of 100 and 200 bar. The Ru/C catalyst was found derived products with the potential to be cofed in crude oil
to present more interesting properties than the classical ones, refineries. A recent report published by BTG (Biomass Tech-
namely concerning the oil yield and deoxygenation level. The nology Group)96 summarizes the advantages and problems related
original acidity of the pyrolysis oil was reduced, as well as its to the several upgrading routes potentially usable in pyrolysis oils
278 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

Figure 1. Fast pyrolysis, hydrotreatment, and hydrocracking combined process. Adapted from Jones et al.94

HDT, when the objective is to arrive at transportation fuels. Bio-oils can be blended with conventional FCC feedstock,
Indeed, many works have been published, but there are many but also dedicated modified FCC units can be envisaged, for
discrepancies and the explanations are scarce. For example, in renewable feedstocks upgrading.
terms of deoxygenation levels, they vary from 10 to 90%, at 3.1. Dedicated FCC for Bio-oils Transformation. A
similar operating conditions and same type of catalysts. dedicated FCC unit for the bio-oils transformation has not
Actually, pyrolysis oils obtained from different sources and become a reality yet. The next sections summarize previous studies
processes are completely different, the hydroprocessing itself is concerning the transformation of model compounds and bio-oils
not well understood, and analysis techniques need further de- over acid catalysts (catalytic cracking).
velopment. As pointed out in the Choudhary and Phillips 3.1.1. Model Compounds Transformation. Depending on
review,41 additional studies should be undertaken in order to the raw-material used to produce bio-oil, the relative amount of
get closer to effective industrial implementation of bio-oils cellulose, hemicellulose, and lignin can vary, and thus the
HDO, namely concerning the effect of bio-oil quality, process corresponding oxygenated building-blocks. This is the reason
parameters, and catalyst optimization. why there are many studies dealing with pure oxygenated
model compounds, in order to understand independently the
3. ACID CATALYSIS UPGRADING OF BIO-OILS key reactions involved in bio-oils transformation. In fact, the
Many studies focusing on bio-oils conversion into liquid trans- transformation of alcohols, aldehydes, ketones, acids, esters,
portation fuels were carried out using acid zeolites as catalysts.1,11 ethers, phenols, furans, and mixtures, over zeolite catalysts has
Renewable feedstock deoxygenation, with solid acid catalysts, can been investigated by several research groups.40,99−108
be achieved at atmospheric pressure and temperatures in the As can be observed in Table 2, alcohols are highly reactive
range of 350−500 °C.1,11,97 The widely used downstream pro- over HZSM-5 catalysts.101 Dehydration is the major and the
cess that can be operated under these conditions is the Fluid first reaction route of alcohols deoxygenation, which can occur
Catalytic Cracking (FCC).98 at relatively low temperatures (around 200 °C), producing
The FCC process plays a very important role in petroleum water and the corresponding alkenes. Once formed, the alkenes
refining, by converting heavy fractions of crude oil into gasoline, oligomerize (250 °C) and, consequently, undergo cyclization
among other valuable products. In this process, there is a con- reactions, yielding aromatics (and/or alkylaromatics) through
tinuous catalyst circulation between the cracking reactor and the hydrogen transfer reactions (>350 °C), which lead to coke
regenerator. The hot catalyst grains promote the endothermic formation.101,104 Ethylene, propylene and butenes are also obtained
hydrocarbon feedstock transformation in the cracking unit, by the cracking of alkyl aromatics and alkanes.104 A second trans-
yielding cracked products and a strongly deactivated catalyst, by formation route should involve the alcohols cracking (decarbon-
rapid coke deposition. In fact, the reactions occurring in the ylation), producing hydrocarbon gases (C1 to C5+), CO, and other
reactor riser quickly lead, after only a few seconds, to an almost alcohols.101,104 It was proposed that the major part of the alcohols
complete deactivation of the catalyst. The spent catalyst particles, are converted into aromatic hydrocarbons on HZSM-5 zeolites,
after separation from the hydrocarbon mixture, are regenerated representing about 40 wt % of the final products.101
by coke combustion, at 650−700 °C. The regenerated hot As for alcohols, aldehydes and ketones deoxygenation over
catalyst particles are then recycled to the riser, so that an almost HZSM-5 zeolites proceeds by dehydration and decarbon-
autothermal operation is achieved. ylation, but they present a lower reactivity (see Table 2).101,105
The FCC catalyst contains acidic HY zeolite as the main In addition, a study concerning the transformation of a model
active phase dispersed into a matrix. Another zeolite, ZSM-5, in bio-oil ketone, the cyclopentanone, over two HY zeolites pre-
its acid form, is commonly used as additive in order to increase senting different acidities, demonstrated that the ketones reac-
the conversion of linear hydrocarbons, leading to the tivity highly depends on the nature of the zeolite acidity. Their
improvement of the gasoline quality by increasing its octane conversion preferentially occurred, at lower temperatures, on
number and, mainly, to an enhancement of the production of the zeolites with lower acid strength, higher Brönsted acid site
valuable light olefins. density, and larger concentration of extraframework aluminum
279 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

species than over the zeolites with higher acid strength but

1.1−1.9; coke, 1.6−4.0; gas, 0.3−2.1; aliphatic, 0.2−0.7; aromatics, 0.3−1.1; eugenol isomers, 47.8−40.4;
lower acid site density.107
The carboxylic acids transformation is very dependent on the
reaction temperature. An important rise of the propanoic acid

3.2−17.8; coke, 6.1−26.7; gas (mainly CO2), 5.3−38.3; aliphatic, 0.7−1.4; aromatics, 0.8−8.0
conversion was observed by increasing the temperature, as
shown in Table 2.101 Carboxylic acids deoxygenation pre-
dominantly occurs by decarboxylation, with CO2 and hydrocarbon
gases formation and, to a lesser extent, by dehydration.101,105
Phenolic species (Table 2) are much less reactive than the
4.7−15.0; coke, 2.1−5.0; gas, 4.8−9.1; aliphatic, 3.2−5.0; aromatics, 29.4−48.5
9.5−12.9; coke, 2.9−7.8; gas, 9.1−15; aliphatic, 2.1−0.2; aromatics, 36.3−39.9

other oxygenated molecules over HZSM-5 zeolite.99,101,104


Alkylated phenols and aromatic hydrocarbons are the main
1.7; coke, 8.3; gas, 42.1; aromatics, 17.6; other furans and phenols, 1.7

reaction products. To a much lower extent, also cracking of the


alkyl groups, leading to the formation of hydrocarbon gases, as
trace; coke, 2.9; gas, trace; aromatics, 55.0; other furans, 13.5
trace; coke, 7.5; gas, 2.9; aromatics, 29.9; other phenols, 57.8

well as condensation-type reactions, leading to coke, occur over


this zeolite. This also showed that decarboxylation and de-
main products: % select

carbonylation were not the major routes for phenols conversion.


Deoxygenation of alkyl- and alkoxyphenols compounds occurred
preferentially through hydrolysis and dehydration (through
hydrogen transfer) reactions.99,103
54.7; coke, 2.0; gas, 22.8; aromatics, 9.4

The conversion of furfural and furan over HZSM-5 zeolite


was also studied,100 being that relatively low conversions were
obtained (Table 2). Furfural catalytic conversion involved de-
carbonylation into furan, through the attack of the zeolite acid
sites. Furan is less reactive than furfural, yielding some poly-
meric materials as a result of the ring-opening reactions, in the
presence of an acid and water. Benzofuran was also found as
one of the major products of furan transformation, resulting
from thermal cycloaddition reactions, involving furan and un-
saturated intermediates originating from the coke.
Hence, the O-compound families presented great differences
in their reactivity, as well as in their reaction pathways. How-
fraction,
fraction,
fraction,
fraction,

fraction,
fraction,
fraction,
fraction,
Table 2. Reactivity and Product Distribution of Pure Oxygenates over HZSM-5100−102

ever, the reactivity of the individual components can change


when mixed, which is due to a synergistic effect established
aqueous
aqueous
aqueous
aqueous

aqueous
aqueous
aqueous
aqueous

between the reactants and the products,101,106 as, for example,


the observed phenols conversion increases in a mixture.
3.1.2. Bio-oils Transformation. The performance of differ-
ent acid catalysts (HZSM-5, HY, HMOR, silicalite and silica−
convn (%)
85.0−98.2

62.1−94.8
24.2−99.9

56.8−60.0

alumina) in the catalytic upgrading of bio-oils was tested by


5.1−9.3

several authors, in a temperature range of 290−450 °C.109−116


65.5
58.9
44.0
100

The products obtained were an organic liquid fraction, con-


sisting of aliphatic and aromatic hydrocarbons and O-compounds,
char, coke, tar or residue, water and gas, essentially consisting of
temp (°C)
330−410

330−410
330−410
330−410
330−410

CO, CO2, and the remaining C1−C6 paraffins and olefins (Table 3).
500

500
500
450

Table 3. Product Distribution, Organic Distillate Fraction


Composition and % Deoxygenation Obtained for the
cat. mass (g)

Upgrading of Bio-oils over Different Catalysts,


WHSV = 3.6 h−1 and 370 °C111,112
3.3
2
10
2
2
2
2
10
10

silica−
HZSM-5 HY HMOR silicalite alumina
WHSV (h−1)

Product Yields (wt %)


char 12.0 21.2 26.6 27.0 16.6
3.6

3.6
3.6
3.6
3.6
4

4
4
1

coke 8.5 14.8 14.8 16.5 24.4


gas 14.3 10.4 14.8 15.4 8.8
organic distillate 33.6 24.6 12.3 10.8 16.8
2-methylcyclopentanone

(ODF)
model compounds
4-methylcyclohexanol

water 28.9 26.2 27.3 24.4 26.8


ODF Composition (wt %)
propanoic acid

aromatic 79.5 16.3 39.4 9.2 5.5


hydrocarbons
methanol

eugenol

furfural
phenol

anisole

aliphatic 3.5 45.0 13.1 36.6 47.8


furan

hydrocarbons
% deoxygenation 90.1 73.3 59.1 73.0 73.3

280 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287


Industrial & Engineering Chemistry Research Review

It was then observed that bio-oils can be upgraded over these of catalyst, the final product being predominantly constituted by
catalysts, with the production of an organic liquid fraction presenting noncyclic aliphatic hydrocarbons. For silicalite, the reaction
high hydrocarbon content. The HZSM-5 zeolite was found to be the pathways are similar to those of acidic zeolites. However, deoxy-
best catalyst tested, with the highest production of liquid products genation by dehydration seems to be lower over silicalite than
and the smaller coke deposition. In fact, a higher deoxygenation rate over acidic zeolites (see Table 3).111,112 It should be noted as
was achieved with the HZSM-5 zeolite (Table 3). well that in the absence of hydrogen, that is, under catalytic crack-
These studies demonstrated that HZSM-5 and HMOR cata- ing conditions, the reduction in the oxygen content is coupled
lysts produced more aromatic hydrocarbons (toluene, xylenes, with a loss in carbon, because part of the oxygen is removed as
and trimethylbenzenes) than HY, silicalite, and silica−alumina, carbon oxides (CO and CO2). Thus, at constant coke + char, a
which were more selective for the production of aliphatic hy- catalyst that favors CO2 formation instead of CO will result in a
drocarbons (alkylated cyclopentene, cyclopropane, pentane and higher hydrocarbon yield. Hence, thermally induced separation
hexane, see Table 3). HY and HMOR could be used to provide and polymerization, as well as thermocatalytic deoxygenation,
hydrocarbons within kerosene boiling point range, whereas cracking, and aromatization, were the main reactions involved in
HZSM-5 is useful to produce hydrocarbons in the gasoline bio-oils conversion. During the bio-oils upgrading, between 30 and
range. Hence, by changing the type of catalyst, it is possible to 40 wt % of the bio-oil was converted into coke.112,113
adjust the composition of the organic distillate fraction resulting Sharma and Bakhshi118 developed a dual reactor system,
from bio-oils deoxygenation. The operating conditions of the where two catalytic reactors, at different temperatures, were
process can also be adjusted to regulate the organic liquid used in series. It was possible to increase nearly 2-fold the organic
fraction and gasoline yields.117 distillate products and the aromatic hydrocarbons yield, as well as
The bio-oils conversion results from thermal and thermo- to reduce the coke plus char formation (10 wt % lower than in the
catalytic transformations (Figure 2).113 Thermal effects induce single reactor). Another studied possibility to reduce coke
deposition was to cofeed steam during bio-oils upgrading over
HZSM-5 zeolites.110 Still, the residue content increased from
negligible values to 8−10 wt %, resulting in a decrease of the
catalyst activity. The presence of steam also induced a change in
the catalyst selectivity: less aromatic and more aliphatic hydro-
carbons were obtained.
As the FCC unit operates in continuous reaction−regeneration
cycles, the behavior of the HZSM-5 zeolite in the upgrading of
bio-oil was tested doing repeated upgrading−regeneration
cycles.119−121 It was observed that the continued regeneration
of the zeolite provokes an irreversible deactivation of the
catalyst, due to a loss of acidity through dealumination, induced
by the presence of steam at high temperatures. To improve the
hydrothermal stability, modified HZSM-5 zeolites have been
tested with higher Si/Al ratios and impregnated with Ni,122
which was revealed to maintain their kinetic behavior for up to
10 reaction−regeneration cycles.
A hydroprocessed oil derived from biomass was also tested in
Figure 2. Scheme of thermal and catalytic transformations of bio-oils the catalytic conversion, in a comparative study with hydro-
over acidic zeolites. Figure adapted with permission from ref 113. processed vacuum gasoil.38 It was observed that the catalytic
Copyright 1995 Elsevier. cracking of hydroprocessed bio-oils gives less gasoline and
gasoil yields, which are compensated by higher coke amounts
the bio-oils separation into two fractions: light and heavy when compared to the hydroprocessed vacuum gasoil. Further-
organics. The thermocatalytic effects are dependent on the more, more light products were obtained for the transformation of
porosity and acidity of the catalyst. Over acid zeolites (i.e., hydrotreated bio-oils. However, the butane and butene yields,
HZSM-5, HY and HMOR, Figure 2), the heavy organic fraction which are more valuable products, were higher for the catalytic
can be cracked into light organics or polymerize to form coke cracking of the conventional hydrotreated vacuum gasoil.
(deposits on the catalyst surface) and char (solid particles not Despite some catalyst stability improvements, a common
deposited on the catalyst). The light organics consist of various concern reported in the literature relative to both HDO and cata-
acids and esters, ketones, alcohols, ethers and phenols. First, lytic conversion processes is still the catalyst deactivation, either by
they are deoxygenated with water and carbon oxides formation. coke deposition and/or chemical changes. In addition, further im-
Then, the hydrocarbons formed undergo cracking and oligo- provements would be important, namely concerning the fuel yields
merization reactions, producing a C2−C6 olefins mixture. These and composition (significant amounts of aromatic compounds).
species can suffer several aromatization reactions (cyclization, 3.2. Coprocessing of Bio-oils in FCC. Biofuels could
alkylation, hydrogen transfer, isomerization), leading to aromatic more quickly relieve the increasing demand for petroleum if
and consequently to coke formation. The results also reveal that economical opportunities for coprocessing them in traditional
gasification of char is a reaction occurring during bio-oil thermo- petroleum refineries are identified and developed. Thus, a
catalytic conversion. The gasification produces carbon oxides and sound option consists of blending the bio-oils with the con-
hydrocarbon gases (mainly olefins). Over silica−alumina cata- ventional FCC refining streams to further processing in the
lysts, the olefins produced are reoxygenated to light organics. existing refinery units. This approach would minimize the
Moreover, cyclization reactions do not take place with this type investment cost by using an already existent infrastructure.
281 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

3.2.1. Coprocessing Studies with Model Compounds. To The greater the size is of the phenolic molecules present in
evaluate the possibility to partially replace the classical FCC the hydrotreated bio-oils, the lower is their effect on the
feedstocks by bio-oils, some recent works analyzed the in- hydrocarbons transformation over the HZSM-5 zeolite, mainly
fluence of the presence of controlled amounts of model oxygenated at higher temperatures.127 In fact, bulky oxygenated molecules
compounds on the cracking properties of the catalysts over (guaiacol) are preferentially adsorbed on the crystals outer
hydrocarbon transformations. Methylcyclohexane and n-heptane surface, from which they are more easily removed at higher
cracking were performed in the presence of phenol or guaiacol temperatures. Contrarily, smaller size phenolic compounds
(oxygenated compounds representative of partially hydrotreated (phenol) are able to enter on the zeolite pores, but their diffu-
bio-oil) over HY and HZSM-5 zeolites (the most active phases in sion is limited due to the polar nature. In this case, the increase
the FCC catalyst) in a fixed-bed reactor, at 350 and 450 °C.123−127 of the temperature does not limit the deactivating effect, the
Phenolic molecules lead to a further deactivation of the zeolites, molecules desorption being kinetically controlled.
along with time-on-stream (TOS), due to their adsorption on The cofeeding of hydrocarbons and oxygenated compounds
the Brönsted and Lewis acid sites, simultaneously to the coke was also tested by Domine et al.128 using a catalytic fixed-bed
molecules deposition from the reactants transformation. This reactor at 530 °C. Acetic acid, acetone, and isopropyl alcohol
effect is of increasing relevance along with TOS, and can be were separately co-injected (2 wt % of the feed) with isooctane,
controlled to avoid important deactivation levels at the first and the mixtures were transformed over an industrial equi-
reaction times, by regulating the O-compound incorporation in librium FCC catalyst. As expected, the addition of the oxygenated
the feed. Since FCC operates at very short residence times (a few compounds causes a decrease of the isooctane conversion. Acetic
seconds), the mentioned cofeeding possibility deserves to be con- acid and isopropyl alcohol led to a stronger initial deactivation of
sidered. Nevertheless, the resistance to the poisons was revealed to the catalyst than acetone. Nevertheless, after successive reaction−
be lower for the HZSM-5 zeolite, because of the narrowness of its regeneration cycles, the deactivation induced by acetone was more
channels. severe than for acetic acid and isopropyl alcohol, for which the
For the large pores HY zeolite,123,125,127 phenol adsorption is deactivation was demonstrated to be somewhat reversible. More-
partially reversible, being thermodynamically controlled: the over, in the presence of the oxygenated compounds, high levels
higher the temperature, the lower the amount of poison ad- of coke were found and coke was not completely removed by
sorbed on the catalyst. The observed detrimental impact of combustion. The O-compounds addition also provoked changes in
phenol on HY activity is significantly dependent on the type of the products distribution. The formation of CO was observed,
hydrocarbon to be transformed by cracking. An important de- resulting from the oxygenated compounds transformation; acetone
activating effect was noticed since the beginning of the reaction apparently suppressed the isooctane secondary cracking reactions,
for the n-heptane transformation (representative of alkane increasing the amounts of isobutane on the effluent.
compounds in conventional FCC charges) with 4% of phenol Few recent works were performed at conditions close to
incorporation in the feed. Contrarily, for methylcyclohexane
those of the FCC process.127,129 The same key bio-oil model O-
(representative of naphthenes in conventional FCC charges),
compounds previously used in HDO studies, representative of
the additional deactivation induced by phenolic molecules, at
the major oxygenated groups (acetic acid, hydroxyacetone,
the same feed composition, was noticed only after a few
phenol or guaiacol) were mixed with a standard gasoil and
minutes of TOS. Furthermore, while for n-heptane the same
tested in a fixed-fluidized bed reactor, using an industrial FCC
deactivation level was observed in the presence of phenol at
equilibrium catalyst (E-CAT) and a mixture of E-CAT and
both studied temperatures, for methylcyclohexane the increase
of the temperature reduced the deactivating effect of phenol. ZSM-5 additive, at 535 °C.127,129 These oxygenated molecules
The impact of the oxygenated molecules is, in fact, related to increased the overall conversion (Figure 3), defined as the fraction
the type of acid sites required for the hydrocarbon cracking:
n-heptane needs stronger acid sites than methylcyclohexane.
Phenols tend to first adsorb on the stronger acid sites, being then
retained on the weaker ones, where the adsorption is less favored
with the increase of the temperature. Hence, the transformation of
the different compound families constituting conventional FCC
feedstocks could be influenced at different levels when copro-
cessed with hydrotreated bio-oils containing phenolic compounds.
The cracking of n-alkanes over HY zeolite seems to be more
affected than naphthenes by the presence of this type of oxygenated
molecules, mainly for the first values of TOS.
On the other hand, n-heptane transformation over HZSM-5
zeolite is less affected by the presence of phenolic molecules
than methylcyclohexane.124,126,127 The n-heptane diffusion
through the narrow and tortuous HZSM-5 channels is more
favored than methylcyclohexane, particularly when the amount Figure 3. Effect of adding different model compounds during cracking
of carbonaceous materials deposited on the zeolite increases. of pure gasoil (■), acetic acid (◊), hydroxyacetone (x), phenol (Δ),
So, the effect of phenol adsorption inside the ZSM-5 pores and guaiacol (○). Figure adapted with permission from refs 127 and
would be stronger over the reduction of methylcyclohexane 129. Copyright 2009/2011 Elsevier.
accessibility to active sites. Considering that linear alkanes
are the main hydrocarbons converted into the ZSM-5 pore of the feed converted into gases, gasoline and coke, since they
network, the impact of the phenolic compounds on this zeolite were considered as products (boiling point in the gasoline
action should not be as important as it could be. range). In spite of this, the gasoil conversion over pure E-CAT
282 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

was not altered by the presence of the O-compounds, while In another study, a heavy hydrotreated bio-oil fraction was
when the ZSM-5 additive was present, a slight decrease of the diluted in LCO and blended with a conventional VGO (vacuum
gasoil conversion was noticed. Under FCC conditions, acetic gas oil), in a proportion of 15/75 wt %.35 This mixture was
acid and hydroxyacetone mainly increased the fuel gas and introduced in a FCC small-scale pilot plant and converted under
liquefied petroleum gas (LPG) fractions. Acetic acid was the FCC operating conditions (520 °C, E-CAT). The results
converted into methane, light olefins (ethylene, propylene and showed that the presence of bio-oil favors the gasoline and diesel
butylenes), CO, and CO2. With hydroxyacetone, an improve- production, but increases the coke yield (about 0.5 wt %). The
ment of the ethylene and propylene production was observed. gasoline produced contained more aromatics and less olefins and
On the other hand, phenol (and guaiacol at high conversions) paraffins. Furthermore, the bio-oils reduced the crackability of
enhanced the gasoline yield (Figure 3). A detailed analysis of the feedstock; that is, for the same catalyst-to-oil ratio, the
the gasoline fractions obtained for the phenol and guaiacol conversion was diminished.
mixtures revealed the presence of methyl- and ethyl-phenols It was also reported that a 20% incorporation level of an
and an increase of the water and benzene on the effluent, as com- hydrotreated bio-oil (oxygen content up to 28 wt %), fed
pared to the pure gasoil transformation. This was also accom- together with 80 wt % crude oil to a MAT FCC unit, produced
panied by an improvement of the LPG yield, in the case of interesting gasoline fractions, equivalent to the one obtained
phenol, and by an increase of the fuel gas, mainly methane, for from pure crude oil.36
guaiacol. Under these operating conditions, phenol can suffer de- A similar bio-oil coprocessing level, namely a mixture of
hydration into benzene and alkylation, and guaiacol can undergo 80 wt % VGO and 20 wt % hydrodeoxygenated bio-oil (oxygen
disproportionation and be converted into phenol, with methane content of 21 wt %), was tested in a fixed-bed reactor
and water formation. Furthermore, it was also observed that the simulating FCC conditions.130 The results revealed that the
ZSM-5 additive effect toward the increase of light compounds bio-oil addition induced an increase of the conversion because
formation is attenuated by these oxygenated molecules, probably the bio-oil already contained molecules in the product range.
because of a preferential interaction of such compounds with the Furthermore, it was noticed that the coke formation tendency
ZSM-5. It was concluded that feed incorporation ratios up to was higher in the case of the coprocessing: at 80% conversion,
12.8 wt % of oxygenated compounds could be processed without the coke yield was 2 wt % higher than with pure VGO cracking.
major problems to the FCC unit. It should be mentioned that The bio-oils incorporation also affected the products distribu-
FCC uses short-contact-time risers, so that the relevant additional tion. An increase of the dry gas yield was also registered, because
catalyst deactivation induced by the oxygenates, observed in the of the improvement of the ethane and methane formation
model reactions studies, shall be the one observed at short TOS. through bio-oils thermal cracking. Furthermore, an enhancement
Nevertheless, the amount of bio-oil incorporated needs to be of the ethylene yield was observed from 35 wt %, for the pure
controlled to keep the ZSM-5 additive effect at an acceptable level, VGO cracking, to 45 wt % in the case of the coprocessing. On
as well as the benzene content in the gasoline range. the other hand, the LPG production decreased for the mixed
3.2.2. FCC Coprocessing Studies with Bio-oils. As feed. The gasoline, LCO, and bottom yields were not signifi-
previously discussed, the direct introduction of bio-oils in the cantly altered by the bio-oils addition to VGO. A detailed
FCC process is limited because of their small miscibility with analysis of the products indicated that the gasoline produced was
hydrocarbons, chemical instability, and high tendency to form richer in branched paraffin and short alkyl-chain benzene deriva-
coke. Therefore, a previous stabilization HDO step of the bio-oils tives, but also contained amounts of phenol and alkylphenols,
should be envisaged before the cofeeding, as it was proposed by which revealed the low reactivity of these oxygenated structures
several authors.34−36,38,130 This step does not need to aim toward under FCC conditions.130,131 Appreciable amounts of CO2 and
a complete deoxygenation, but only to the reduction of the water were also found in the effluent, resulting from deoxygenation
highly reactive components that lead to coke formation.36 reactions, decarboxylation, and dehydration, respectively.130,131 A
Mixtures up to 20 wt % of hydrotreated bio-oils and con- final product composition poor in H2 and containing more coke,
ventional FCC streams were coprocessed over equilibrium FCC aromatics, and olefins was obtained through the coprocessing,
catalysts, in order to analyze the effects of partially replacing the because the oxygen removal from hydrotreated bio-oils occurred
FCC feedstocks by bio-oils on the conversion, products preferentially by dehydration.130,131
distribution, and coke formation. Thus, available data, as shown in Table 4, suggest that
A coprocessing of 15 wt % of hydrotreated bio-oil and LCO controlled cofeeding of bio-oils toward incorporation levels of
(light cycle oil) was performed in a MAT fixed bed reactor, about 20%, together with a conventional FCC feedstock, is not
using two commercially available catalysts (RE-USY), with that critical, seeming to be a technically viable option. Despite
different rare-earth contents.34 The catalyst with an RE2O3 amount some changes in products selectivities reported, the quality of
closer to that generally used for a maximum gasoline yield (0.5 wt %) the produced gasoline was generally preserved (Table 5).
was found to be the more suitable for this application. This catalyst
preserved the conversion of heavy compounds within a limited level,
producing acceptable levels of coke (<1 wt %) and gasoline yields
4. FINAL REMARKS
(23−25 wt %). Moreover, it produced a higher RON (research The introduction of renewable materials, namely nonedible
octane number) gasoline due to the greater aromatic content. biomass derived oils (bio-oils), into downstream operations is a
Nevertheless, a lower amount of branched paraffin and saturated timely topic. Two main refining processes have been envisaged
naphthenes were obtained when compared to catalysts with larger to upgrade these bio-oils toward the production of useful liquid
pores. Comparing the biogasoline produced with a common FCC fuels (second generation biofuels): HDO and catalytic cracking
gasoline, it was verified that olefin and naphthene yields are lower, conversion. Many nonedible feedstocks can be used, but they
whereas the amount of aromatic compounds increases. However, the all have in common the presence of oxygenated compounds,
obtained biogasoline still met the environmental gasoline specifica- which have to be (at least partially) transformed, by oxygen
tions. removal, in order to allow blending with hydrocarbon feedstocks.
283 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research Review

Table 4. Comparison of Products Yields Obtained for catalysts. Supported Ru, Rh, and Pt appeared to be potential
Classical and Bio-oil Blended Feedstocks under FCC catalysts for HDO, but it should be also taken into con-
Conditions.35,36,130 sideration that the possible processing costs rise when noble
convn dry gas LPG gasoline LCO coke
metals are used in a large scale process. Less expensive metal
(%) feed (wt %) (wt %) (wt %) (wt %) (wt %) catalysts, such as Ni and Cu, are potential alternatives, but
65 VGO + 15%LCO 42 19 3.5 further work on these systems is necessary.
VGO + 15% 43 20 4 HDO of bio-oils has been also reported to be accomplished
(LCO+bio-oil) over classical hydrotreatment catalysts, as well as over noble
60 VGO 1.5 8.5 44 25 5.9 metals. In fact, noble metal catalysts are known to be active in
VGO + 20% bio- 2.5 11 44 23 7.8 hydrogenation and hydrodeoxygenation. The particular case of
oil (28% O)
80 VGO 2 26 47 19 3.8
Ru presents high activity toward the reduction of a wide range
VGO + 20% bio- 2.5 24 46 19 5.8 of different oxygenates. The development of bio-oils HDO
oil (21% O) studies and its integration in the refineries are, as well, being
carried out by several companies and researchers, but the costs
related with capital and hydrogen consumption are still very
Table 5. Effect of Bio-oil Incorporation on FCC Gasoline
high. High costs and the need for further catalysts optimization,
Composition and RON.34
among other factors, are thus preventing implementation of
compounds contents (wt %) LCO LCO + 15% bio-oil dedicated HDO units. The immediate and easy coprocessing of
n-paraffins 2.5 2.3 bio-oils with classical crude fractions, in the existing HDT
iso-paraffins 16.7 16.5 processes, does not seem to be feasible. The optimal perfor-
naphthenes 5.2 5.4 mance of current industrial HDT catalysts for petroleum
n-olefins 6.2 5.4 feedstocks cannot be reproduced over bio-oils HDO.
iso-olefins 10.3 8.8 The cracking of bio-oils with solid acid catalysts is less well
diolefins 0.7 0.5 developed than hydrotreating, but a wide range of possibilities
aromatics 58.4 61.1 have been recently explored, with mono- and multifunctional
RON 94.7 95.8
zeolite catalysts, revealing that this is an interesting alternative
for bio-oils transformation into liquid fuels. Compared to
The comparison of both technologies, concerning the main HDO, it is a cheaper alternative, offering significant processing
interesting features and drawbacks of each one on the bio-oils and economic advantages, because hydrogen is not required
upgrading, is presented in Table 6. and operation is under atmospheric pressure.
However, the existence of a dedicated FCC unit for the
Table 6. Comparison of HDO and FCC Technologies for biomass renewable feedstocks transformation needs further
Bio-oils Upgrading development. The poor hydrocarbons yields, the extensive
technology interesting features drawbacks coking observed during catalytic cracking of bio-oils, and the
HDO quality of the fuels obtained (significant amounts of phenolic
dedicated recent improvements on hydrogen consumption compounds) are important reported drawbacks that need to be
HDO catalyst performances overcome.
no need of catalyst H2S catalysts costs Taking into account the actual FCC process features, as in
pretreatments
HDT co- lower capital costs HDN and HDS efficiency
the Resid FCC Process, where the extensive coke deposition is
processing significantly affected by the a solved technological problem, and considering some recent
oxygenates studies, it seems that the coprocessing of bio-oils, in com-
FCC bination with conventional feedstocks, in existing FCC units, is
dedicated cheaper than HDO extensive coking
FCC the most promising bio-oils upgrading alternative in a short-
quality of the fuels term. In addition, relative small investment costs are involved.
FCC co- irrelevant investment need to control of bio-oils Nevertheless, the extent of bio-oil incorporation in the FCC
processing costs incorporation in the FCC feed feed should be controlled to remain at previously studied levels,
good quality gasoline HDO previous stabilization step so that the FCC outcoming products can maintain the desired
specifications
quality, not exceeding the aromatics content specification in
gasoline and keeping at acceptable levels the octane number of
For HDO option, the choice of a catalyst plays a crucial role, gasoline and the light olefins in LPG. This will be a function of
not only in the extent of oxygen removal, but also on the
the bio-oil properties, which depend on the severity of the
predominant pathway, that is, through dehydrogenation,
decarbonylation, or decarboxylation. In a nutshell, the pathway previous partial (and less expensive) HDO upgrading step.
of choice relays on the feedstock availability and logistics, Other aspects needing further research developments, con-
catalyst cost and hydrogen consumption. A great amount of cerning the coprocessing option, are related to the regener-
research work has been devoted to the model O-compounds ability of the catalysts, which will be submitted to numerous
transformations through HDO. Conventional hydrotreating reaction−regeneration cycles in the FCC process, and that was
catalysts were extensively tested in these studies. Nevertheless, proven to be excellent in the pure crude oil FCC. At least, such
catalytic systems based on noble metals supported on ZrO2, as results are expected to be more promising than those obtained
well as transition metal phosphides on silica, showed, in some with a full and nontreated bio-oil feedstock transformation
recent studies, to be more stable than classical CoMo/Al2O3 reported in the literature.
284 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287
Industrial & Engineering Chemistry Research


Review

AUTHOR INFORMATION (25) Venderbosch, R. H.; Ardiyanti, A. R.; Wildschut, J.; Oasmaa, A.
Stabilization of biomass-derived pyrolysis oils. J. Chem. Technol.
Corresponding Author Biotechnol. 2010, 85, 674.
*Tel.: +351218419073. Fax: +351218419198. E-mail: jmlopes@ (26) Zhang, Q.; Chang, J.; Wang, T.; Xu, Y. Review of biomass
ist.utl.pt. pyrolysis oil properties and upgrading research. Energy Convs. Manage.
2007, 48, 87.
Notes
(27) Kersten, S. R. A.; van Swaaij, W. P. M.; Lefferts, L.; Seshan, K.
The authors declare no competing financial interest. Options for catalysis in thermochemical conversion of biomass into

■ REFERENCES
(1) Huber, G. W.; Iborra, S.; Corma, A. Synthesis of transportation
fuels. In: Catalysis for Renewables: From feedstock to energy production;
Centi, G., van Santen, R. A., Ed.; Wiley-VCH: Weinheim, 2007; p 119.
(28) Corma, A.; Huber, G. W.; Sauvanaud, L.; O’Connor, P.
Processing biomass-derived oxygenates in the oil refinery: Catalytic
fuels from biomass: Chemistry, catalysts, and engineering. Chem. Rev.
cracking (FCC) reaction pathways and role of catalyst. J. Catal. 2007,
2006, 106, 4044.
247, 307.
(2) McKendry, P. Energy production from biomass (part 1):
(29) Elliot, D. C.; Hart, T. R.; Neuenschwander, G. G.; Rotness, L. J.;
Overview of biomass. Bioresour. Technol. 2002, 83, 37.
Zacher, A. H. Catalytic hydroprocessing of biomass fast pyrolysis oil to
(3) Lange, J.-P. Lignocellulose conversion: An introduction to
chemistry, process and economics. Biofuels Bioprod. Biorefin. 2007, 1, produce hydrocarbon products. Environ. Progress Sust. Energy 2009, 28,
39. 441.
(4) Bridgwater, T. Applications for utilisation of liquids produced by (30) de Wild, P.; der Laan, R. V.; Kloekhorst, A.; Heeres, E. Lignin
fast pyrolysis of biomass. Biomass Bioenergy 2007, 31, I. valorisation for chemicals and (transportation) fuels via (catalytic)
(5) Demirbas, M. F.; Balat, M.; Balat, H. Potential contribution of pyrolysis and hydroxydeoxygenation. Environ. Progress Sust. Energy
biomass to the sustainable energy development. Energy Convs. Manage. 2009, 28, 461.
2009, 50, 1746. (31) Laurent, E.; Delmon, B. Study of the hydrodeoxygenation of
(6) Escobar, J. C.; Lora, E. S.; Venturini, O. J.; Yáñez, E. E.; Castillo, carbonyl, carboxyl, and guaiacyl groups over sulfided CoMo/γ-Al2O3
E. F.; Almazam, O. Biofuels: Environment, technology and food and NiMo/γ-Al2O3 catalysts. I. Catalytic reaction schemes. Appl. Catal.
security. Renew. Sustain. Energy Rev. 2009, 13, 1275. A: Gen. 1994, 109, 77.
(7) Stéphane, O.; Daniel, T. Biorefinery: Toward an industrial (32) Centeno, A.; Laurent, E.; Delmon, B. Influence of the support of
metabolism. Biochimie 2009, 91, 659. CoMo sulphide catalysts and of the addition of potassium and
(8) Stö cker, M. Biofuels and biomass-to-liquid fuels in the platinum on the catalytic performances for the hydroxydeoxygenation
biorefinery: Catalytic conversion of lignocellulosic biomass using of carbonyl, carboxyl, and guaiacol-type molecules. J. Catal. 1995, 154,
porous materials. Angew. Chem., Int. Ed. 2008, 47, 9200. 288.
(9) Mohan, D.; Pittman, C. U., Jr.; Steele, P. H. Pyrolysis of wood/ (33) Bui, V. N.; Toussaint, G.; Laurenti, D.; Mirodatos, C.; Geantet,
biomass for bio-oil: A critical review. Energy Fuels 2006, 20, 848. C. Co-processing of pyrolysis bio oils and gas oil for new generation of
(10) Elliott, D. C.; Beckman, D.; Bridgewater, A. V.; Diebold, J. P.; biofuels: Hydrodeoxygenation of guaiacol and SRGO mixed feed.
Gevert, S. B.; Solantausta, Y. Developments in thermal liquefaction of Catal. Today 2009, 143, 172.
biomass: 1983−1990. Energy Fuels 1991, 5, 399. (34) Samolada, M. C.; Baldauf, W.; Vasalos, I. A. Production of a bio-
(11) Bridgwater, A. V. Catalysis in thermal biomass conversion. Appl. gasoline by upgrading biomass flash pyrolysis liquids via hydrogen
Catal. A: Gen. 1994, 116, 5. processing and catalytic cracking. Fuel 1998, 77, 1667.
(12) Demirbaş, A. Mechanisms of liquefaction and pyrolysis reactions (35) Lappas, A. A.; Bezergianni, S.; Vasalos, I. A. Production of
of biomass. Energy Convs. Manage. 2000, 41, 633. biofuels via co-processing in conventional refining processes. Catal.
(13) Demirbaş, A. Biomass resource facilities and biomass conversion Today 2009, 145, 55.
processing for fuels and chemicals. Energy Convs. Manage. 2001, 42, (36) Mercader, F. M.; Groeneveld, M. J.; Kersten, S. R. A.; Way, N.
1357. W. J.; Schaverien, C. J.; Hogendoorn, J. A. Production of advanced
(14) Bridgwater, A. V.; Meier, D.; Radlein, D. An overview of fast biofuels: Co-processing of upgraded pyrolysis oil in standard refinery
pyrolysis of biomass. Org. Geochem. 1999, 30, 1479. units. Appl. Catal. B: Environ. 2010, 96, 57.
(15) Venderbosch, R. H.; Prins, W. Fast pyrolysis technology (37) Darmstadt, H.; Garcia-Perez, M.; Adnot, A.; Chaala, A.;
development. Biofuels Bioprod. Biorefin. 2010, 4, 178. Kretschmer, D.; Roy, C. Corrosion of metals by bio-oil obtained by
(16) Qi, Z.; Jie, C.; Tiejun, W.; Ying, X. Review of biomass pyrolysis vacuum pyrolysis of softwood bark residues. An X-ray photoelectron
oil properties and upgrading research. Energy Convs. Manage. 2007, 48, spectroscopy and Auger electron spectroscopy study. Energy Fuels
87. 2004, 18, 1291.
(17) Qiang, L.; Wen-Zhi, L.; Xi-Feng, Z. Overview of fuel properties (38) Gevert, B. S.; Otterstedt, J.-E. Upgrading of directly liquefied
of biomass fast pyrolysis oils. Energy Convs. Manage. 2009, 50, 1376. biomass to transportation fuels: catalytic cracking. Biomass 1987, 14,
(18) Oasmaa, A.; Czernik, S. Fuel oil quality of biomass pyrolysis oils: 173.
State of the art for the end users. Energy Fuels 1999, 13, 914. (39) Alonso, D. M.; Bond, J. Q.; Dumesic, J. A. Catalytic conversion
(19) Czernik, S.; Bridgwater, A. V. Overview of applications of of biomass to fuels. Green Chem. 2010, 12, 1493.
biomass fast pyrolysis oil. Energy Fuels 2004, 18, 590. (40) Mortensen, P. M.; Grunwaldt, J.-D.; Jensen, P. A.; Knudsen, K.
(20) Elliott, D. C. Historical developments in hydroprocessing bio- G.; Jensen, A. D. A review of catalytic upgrading of bio-oil to engine
oils. Energy Fuels 2007, 21, 1792. fuels. Appl. Catal. A: Gen. 2011, 407, 1.
(21) Huber, G. W.; Corma, A. Synergies between Bio- and oil (41) Choudhary, T. V.; Phillips, C. B. Renewable fuels via catalytic
refineries for the production of fuels from biomass. Angew. Chem. hydrodeoxygenation. Appl. Catal. A: Gen. 2011, 397, 1.
2007, 46, 7184. (42) Bridgwater, A. V. Review of fast pyrolysis of biomass and
(22) Furimsky, E. Catalytic hydrodeoxygenation. Appl. Catal. A: Gen. product upgrading. Biomass Bioenergy 2012, 38, 68.
2000, 199, 147. (43) Furimsky, E. Chemistry of catalytic hydrodeoxygenation. Catal.
(23) Gandarias, I.; Barrio, V. L.; Requies, J.; Á rias, P. L.; Cambra, J. Rev.-Sci. Eng. 1983, 25, 421.
F.; Güemez, M. B. From biomass to fuels: Hydrotreating of (44) Gutierrez, A.; Kaila, R. K.; Honkela, M. L.; Slioor, R.; Krause, A.
oxygenated compounds. Int. J. Hydrogen Energy 2008, 33, 3485. O. I. Hydrodeoxygenation of guaiacol on noble metal catalysts. Catal.
(24) Maggi, R.; Delmon, B. A review of catalytic hydrotreating Today 2009, 147, 239.
processes for the upgrading of liquids produced by flash pyrolysis. (45) Zhao, H. Y.; Li, D.; Bui, P.; Oyama, S. T. Hydrodeoxygenation
Stud. Surf. Sci. Catal. 1997, 106, 99. of guaiacol as model compound for pyrolysis oil on transition metal

285 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287


Industrial & Engineering Chemistry Research Review

phosphide hydroprocessing catalysts. Appl. Catal. A: Gen. 2010, 391, (67) Şenol, O.I.̇ ; Viljava, T.-R.; Krause, A. O. I. Effect of sulphiding
305. agents on the hydrodeoxygenation of aliphatic esters on sulphided
(46) Popov, A.; Kondratieva, E.; Goupil, J. M.; Mariey, L.; Bazin, P.; catalysts. Appl. Catal. A: Gen. 2007, 326, 236.
Gilson, J.-P.; Travert, A.; Maugé, F. Bio-oils hydrodeoxygenation: (68) Snåre, M.; Kubičková, I.; Mäki-Arvela, P.; Eränen, K.; Murzin,
Adsorption of phenolic molecules on oxidic catalyst supports. J. Phys. D. Y. Heterogeneous catalytic deoxygenation of stearic acid for
Chem. 2010, 114, 15661. production of biodiesel. Ind. Eng. Chem. Res. 2006, 45, 5708.
(47) Vuori, A.; Helenius, A.; Bredenberg, J.B.-S. Influence of sulphur (69) Şenol, O.I.̇ ; Ryymin, E.-M.; Viljava, T.-R.; Krause, A. O. I.
level on hydrodeoxygenation. Appl. Catal. 1989, 52, 41. Reactions of methyl heptanoate hydrodeoxygenation on sulphided
(48) Huuska, M.; Rintala, J. Effect of catalyst acidity on the catalysts. J. Mol. Catal. A 2007, 268, 1.
hydrogenolysis of anisole. J. Catal. 1985, 94, 230. (70) Katrizky, A. R.; Rees, C. W. Comprehensive Heterocyclic Chemistry
(49) Yakovlev, V. A.; Khromova, S. A.; Sherstyuk, O. V.; Dundich, V. Vol. 4; Pergamon Press: New York, 1984.
O.; Ermakov, D. Y.; Novopashina, V. M.; Lebedev, M. Y.; (71) Chary, K. V. R.; Rama Rao, K. S.; Muralidhar, G.; Kanta Rao, P.
Bulavchenko, O.; Parmon, V. N. Development of new catalytic Hydrodeoxygenation of furan by carbon supported molybdenum
systems for upgraded biofuels production from bio-crude-oil and sulphide catalysts. Carbon 1991, 29, 478.
biodiesel. Catal. Today 2009, 144, 362. (72) Kordulis, C.; Gouromihou, A.; Lycourghiotis, A.; Papadopoulo,
(50) Shin, E.-J.; Keane, M. A. Gas-phase hydrogenation/hydro- C.; Matralis, H. K. Fluorinated hydrotreatment catalysts; Hydro-
genolysis of phenol over supported nickel catalysts. Ind. Eng. Chem. deoxygenation and hydrocracking on fluorine−nickel−molybdenum/
Res. 2000, 39, 883. γ-alumina catalysts. Appl. Catal. 1990, 67, 39.
(51) Yang, Y.; Gilbert, A.; Xu, C. Hydrodeoxygenation of bio-crude (73) Furimsky, E. Deactivation of molibdate catalyst during
in supercritical hexane with sulfide CoMo and CoMoP catalysts hydrodeoxygenation of tetrahydrofuran. Ind. Eng. Chem. Prod. Res.
supported on MgO: A model study using phenol. Appl. Catal. A: Gen. Dev. 1983, 22, 34.
2009, 360, 242. (74) Lee, C.-L.; Ollis, D. F. Catalytic hydrodeoxygenation of
(52) Shin, E.-J.; Keane, M. A. Catalytic hydrogen treatment of benzofuran and o-ethylphenol. J. Catal. 1984, 87, 325.
aromatic alcohols. J. Catal. 1998, 173, 450. (75) La Vopa, V.; Satterfield, C. N. Catalytic hydrodeoxygenation of
(53) Ahmad, M. M.; Nordin, M. F. R.; Azizan, M. T. Upgrading bio- dibenzofuran. Energy Fuels 1987, 1, 323.
oil into high value hydrocarbons via hydroxydeoxygenation. Am. J. (76) Kreuzer, K.; Kramer, R. Support effects in the hydrogenolysis of
Appl. Sci. 2010, 7, 746. tetrahydrofuran on platinum catalysts. J. Catal. 1997, 167, 391.
(54) Wang, W.; Yang, Y.; Luo, H.; Liu, W. Characterization and (77) Furimsky, E. The mechanism of catalytic hydrodeoxygenation of
hydrodeoxygenation properties of Co promoted Ni−Mo−B amor- furan. Appl. Catal. 1983, 6, 159.
phous catalysts: Influence of Co content. React. Kinet. Mech. Catal. (78) Edelman, M. C.; Maholand, M. K.; Baldwin, R. M.; Cowley, S.
2010, 101, 105. W. Vapour-phase catalytic hydrodeoxygenation of benzofuran. J. Catal.
(55) Massoth, F. E.; Politzer, P.; Concha, M. C.; Murray, J. S.; 1988, 111, 243.
Jakowski, J.; Simons, J. Catalytic hydrodeoxygenation of methyl- (79) Bartok, M.; Szollosi, G.; Apjok, J. Mechanism of hydrogenolysis
substituted phenols: Correlations of kinetic parameters with molecular and isomerisation of oxacycloalkanes on metals, XVI. Transformation
properties. J. Phys. Chem. B 2006, 110, 14283. of tetrahydrofuran on platinum catalysts. React. Kinet. Catal. Lett. 1998,
(56) Yang, Y. Q.; Tye, C. T.; Smith, K. J. Influence of MoS2 catalyst 64, 21.
morphology on the hydrodeoxygenation of phenols. Catal. Commun. (80) Bunch, A. Y.; Wang, X.; Ozkan, U. S. Adsorption characteristics
2008, 9, 1364. of reduced Mo and Ni−Mo catalysts in the hydrodeoxygenation of
(57) Laurent, E.; Delmon, B. Influence of oxygen-, nitrogen-, and benzofuran. Appl. Catal. A: Gen. 2008, 346, 96.
sulfur-containing compounds on the hydrodeoxygenation of phenols (81) Goudriaan, F.; Peferoen, D. G. R. Liquids fuels from biomass via
over sulfided CoMo/γ-Al2O3 and NiMo/ γ-Al2O3 catalysts. Ind. Eng. a hydrothermal process. Chem. Eng. Sci. 1990, 45, 2729.
Chem. Res. 1993, 32, 2516. (82) Klass, D. L. Energy from Biomass and Wastes XI; IGT: Chicago,
(58) Wandas, R.; Surygala, J.; Sliwka, E. Conversion of cresols and 1988.
naphtalene in the hydroprocessing of three-component model (83) Bridgwater, A. V. Advances in Thermochemical Biomass
mixtures simulating fast pyrolysis tars. Fuel 1996, 75, 687. Conversion; Blackie: London, 1992.
(59) Odebunmi, E. O.; Ollis, D. F. Catalytic hydrodeoxygenation. 1. (84) Elliot, D. C., Baker, E. G. Energy from Biomass and Wastes X;
Conversions of o-, p-, and m-cresols. J. Catal. 1983, 80, 56. Elsevier: New York, 1986.
(60) Weigold, H. Behaviour of Co−Mo−Al2O3 catalysts in the (85) Baker, E. G.; Elliot, D. C. In Research in Thermochemical Biomass
hydrodeoxygenation of phenols. Fuel 1982, 61, 1021. Conversion; Bridgwater, A. V., Kuster, J. L., Eds., Elsevier: London,
(61) Gevert, B. S.; Otterstedt, J.-E.; Massoth, F. E. Kinetics of the 1988; p 883.
HDO of methyl-substituted phenols. Appl. Catal. 1987, 31, 119. (86) Bredenberg, J.B.-son; Huusta, M.; Räty, J.; Korpio, M.
(62) Mori, A.; Mizusaki, T.; Ikawa, T.; Maegawa, T.; Monguchi, Y.; Hydrogenolysis and hydrocracking of the carbon-oxygen bond: I.
Sajiki, H. Palladium on carbon-diethylamine-mediated hydrodeoxyge- Hydrocracking of some simple aromatic O-compounds. J. Catal. 1982,
nation of phenol derivatives under mild conditions. Tetrahedron 2007, 77, 242.
63, 1270. (87) Bredenberg, J.B.-son; Huusta, M.; Toropainen, P. Hydro-
(63) Bejblová, M.; Zámostný, P.; Č ervený, L.; Č ejka, J. Hydro- genolysis of differently substituted methoxyphenols. J. Catal. 1989,
deoxygenation of benzophenone on Pd catalysts. Appl. Catal. A: Gen. 120, 401.
2005, 296, 169. (88) Kallury, R. K. M. R.; Restivo, W. M.; Tidwell, T. T.; Boocock, D.
(64) Procházková, D.; Zámostný, P.; Bejblová, M.; Č ervený, L.; G. B.; Crimi, A.; Douglas, J. Hydrodeoxygenation of hydroxyl,
Č ejka, J. Hydrodeoxygenation of aldehydes catalyzed by supported methoxy and methyl phenols with molybdenum oxide/nickel oxide/
palladium catalysts. Appl. Catal. A: Gen. 2007, 332, 56. alumina catalyst. J. Catal. 1985, 96, 535.
(65) Pham, T. T.; Lobban, L. L.; Resasco, D.; Mallinson, R. G. (89) French, R. J.; Hrdlicka, J.; Baldwin, J. R. Mild hydrotreating of
Hydrogenation and hydrodeoxygenation of 2-methyl-2-pentanal on biomass pyrolysis oils to produce a suitable refinery feedstock. Environ.
supported metal catalysts. J. Catal. 2009, 266, 9. Progress Sustain. Energy 2010, 29, 142.
(66) Şenol, O.I.̇ ; Viljava, T.-R.; Krause, A. O. I. Hydrodeoxygenation (90) Churin, E.; Maggi, R.; Grange, P.; Delmon, B. In Research in
of methyl esters on sulfide NiMo/γ-Al2O3 and CoMo/γ-Al2O3 Thermochemical Biomass Conversion; Bridgwater, A. V., Kuster, J. L.,
catalysts. Catal. Today 2005, 100, 331. Eds.; Elsevier: London, 1988; p 896.

286 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287


Industrial & Engineering Chemistry Research Review

(91) Rocha, J. D.; Luengo, C. A.; Snape, C. E. Hydrodeoxygenation (113) Adjaye, J. D.; Bakhshi, N. N. Production of hydrocarbons by
of oils from cellulose in single and two-stage hydropyrolysis. Renew. catalytic upgrading of a fast pyrolysis bio-oil. Part II: Comparative
Energy 1996, 9, 950. catalyst performance and reaction pathways. Fuel Process. Technol.
(92) Wildschut, J.; Mahfud, F. H.; Venderbosch, R. H.; Heeres, H. J. 1995, 45, 185.
Hydrotreatment of fast pyrolysis oil using heterogeneous noble-metal (114) Williams, P. T.; Horne, P. A. The influence of catalyst type on
catalysts. Ind. Eng. Chem. Res. 2009, 48, 10324. the composition of upgraded biomass pyrolysis oils. J. Anal. Appl.
(93) McCall, M. J.; Brandvold, T. A. Fuel and fuel blending Pyrol. 1995, 31, 39.
components from biomass derived pyrolysis oil. US Patent 2009/ (115) Adjaye, J. D.; Katikaneni, S. P. R.; Bakhshi, N. N. Catalytic
0253948, 2009. conversion of a biofuel to hydrocarbons: Effect of mixtures of HZSM-5
(94) Jones, S. B.; Holladay, J. E.; Valkenburg, C.; Stevens, D. J.; and silica-alumina catalysts on product distribution. Fuel Process.
Walton, C. W.; Kinchin, C.; Elliott, D. C.; Czernik, S. Production of Technol. 1996, 48, 115.
gasoline and diesel from biomass via fast pyrolysis; hydrotreating and (116) Vitolo, S.; Seggiani, M.; Frediani, P.; Ambrosini, G.; Politi, L.
hydrocracking: A design case. Report 18284PNNL; Pacific Northwest Catalytic upgrading of pyrolytic oil to fuels over different zeolites. Fuel
National Laboratory: Richland, WA, 2009. 1999, 78, 1147.
(95) Wright, M. M.; Satrio, J. A.; Brown, R. C.; Daugaard, D. E.; Hsu, (117) Hew, K. L.; Tamidi, A. M.; Yusup, S.; Lee, K. T.; Ahmad, M.
D. D. Techno-economic analysis of biomass fast pyrolysis to M. Catalytic cracking of bio-oil to organic liquid product (OLP).
transportation fuels. Technical Report NREL/TP-6A20-46586; National Bioresour. Technol. 2010, 101, 8855.
Renewable Energy Laboratory: Golden CO, November 2010. (118) Sharma, R. K.; Bakhshi, N. N. Catalytic conversion of fast
(96) Venderbosch, R. H.; Ardiyanti, A.; Wildschut, J.; Oasmaa, A.; pyrolysis oil to hydrocarbon fuels over HZSM-5 in a dual reactor
Heeres, H. J. Insights in the hydroprocessing of biomass derived pyrolysis system. Biomass Bioenergy 1993, 5, 445.
oils; Paper Biocoup, July 2009; BTG Biomass Technology Group: The (119) Vitolo, S.; Bresci, B.; Seggiani, M.; Gallo, M. G. Catalytic
Netherlands, 2009. upgrading of pyrolytic oils over HZSM-5 zeolite: Behaviour of the
(97) Bridgwater, A. V. Production of high grade fuels and chemicals catalyst when used in repeated-regeneration cycles. Fuel 2001, 80, 17.
from catalytic pyrolysis of biomass. Catal. Today 1996, 29, 285. (120) Gayubo, A. G.; Aguayo, T.; Atutxa, A.; Prieto, R.; Bilbao, J.
(98) Sadeghbeigi, R. Fluid Catalytic Cracking HandbookDesign, Deactivation of a HZSM-5 zeolite catalyst in the transformation of the
Operation and Troubleshooting of FCC Facilities, 2nd ed.; Gulf aqueous fraction of biomass pyrolysis oil into hydrocarbons. Energy
Publishing Company: USA, 2000. Fuels 2004, 18, 1640.
(99) Chantal, P. D.; Kaliaguine, S.; Grandmaison, J. L. Reactions of (121) Guo, X.; Zheng, Y.; Zhang, B.; Chen, J. Analysis of coke
phenolic compounds over HZSM-5. Appl. Catal. 1985, 18, 133. precursor on catalyst and study on regeneration of catalyst in
(100) Grandmaison, J. L.; Chantal, P. D.; Kaliaguine, S. C. upgrading bio-oil. Biomass Bioenergy 2009, 33, 1469.
Conversion of furanic compounds over H-ZSM-5 zeolite. Fuel 1990, (122) Valle, B.; Gayubo, A. G.; Alonso, A.; Aguayo, A. T.; Bilbao, J.
69, 1058. Hydrothermally stable HZSM-5 zeolite catalysts for the transformation
(101) Adjaye, J. D.; Bakhshi, N. N. Catalytic conversion of a biomass- of crude bio-oil into hydrocarbons. Appl. Catal. B: Environ. 2010, 100,
derived oil to fuels and chemicals I: Model compound studies and 318.
reaction pathways. Biomass Bioenergy 1995, 8, 131. (123) Graça, I.; Comparot, J.-D.; Laforge, S.; Magnoux, P.; Lopes, J.
(102) Horne, P. A.; Williams, P. T. Reaction of oxygenated biomass M.; Ribeiro, M. F.; Ramôa Ribeiro, F. Effect of phenol addition on the
pyrolysis model compounds over a ZSM-5 catalyst. Renew. Energy performances of H-Y zeolite during methylcyclohexane transforma-
1996, 7, 131. tion. Appl. Catalysis A: Gen. 2009, 353, 123.
(103) Samolada, M. C.; Papafotica, A.; Vasalos, I. A. Catalyst (124) Graça, I.; Comparot, J.-D.; Laforge, S.; Magnoux, P.; Lopes, J.
evaluation for catalytic biomass pyrolysis. Energy Fuels 2000, 14, 1161. M.; Ribeiro, M. F.; Ramôa Ribeiro, F. Influence of phenol addition on
(104) Gayubo, A. G.; Aguayo, A. T.; Atutxa, A.; Aguado, R.; Bilbao, J. the H-ZSM-5 zeolite catalytic properties during methylcyclohexane
Transformation of oxygenate components of biomass pyrolysis oil on a transformation. Energy Fuels 2009, 23, 4224.
HZSM-5 zeolite. I. Alcohols and Phenols. Ind. Eng. Chem. Res. 2004, (125) Graça, I.; Lopes, J. M.; Ribeiro, M. F.; Laforge, S.; Magnoux, P.;
43, 2610. Ramôa Ribeiro, F. Effect of phenol adsorption on HY zeolite for n-
(105) Gayubo, A. G.; Aguayo, A. T.; Atutxa, A.; Aguado, R.; Olazar, heptane cracking: Comparison with methylcyclohexane. Appl. Catal. A:
M.; Bilbao, J. Transformation of oxygenate components of biomass Gen. 2010, 385, 178.
pyrolysis oil on a HZSM-5 zeolite. II. Aldehydes, ketones, and acids. (126) Graça, I.; Lopes, J. M.; Ribeiro, M. F.; Laforge, S.; Magnoux, P.;
Ind. Eng. Chem. Res. 2004, 43, 2619. Ramôa Ribeiro, F. Bio-oils and FCC feedstocks co-processing: Impact
(106) Gayubo, A. G.; Aguayo, A. T.; Atutxa, A.; Valle, B.; Bilbao, J. of phenolic molecules on FCC hydrocarbons transformation over
Undesired components in the transformation of biomass pyrolysis oil MFI. Fuel 2011, 90, 467.
into hydrocarbons on an HZSM-5 zeolite catalyst. J. Chem. Technol. (127) Graça, I.; Lopes, J. M.; Ribeiro, M. F.; Ramôa Ribeiro, F.;
Cerqueira, H. S.; Almeida, M. B. B. Catalytic cracking in the presence
Biotechnol. 2005, 80, 1244.
of guaiacol. Appl. Catal. B: Environ. 2011, 101, 613.
(107) Huang, J.; Long, W.; Agrawal, P. K.; Jones, C. Effects of acidity
(128) Domine, M. E.; van Veen, A. C.; Schuurman, Y.; Mirodatos, C.
on the conversion of the model bio-oil ketone cyclopentanone on H-Y
Coprocessing of oxygenated biomass compounds and hydrocarbons
zeolites. J. Phys. Chem. C 2009, 113, 16702.
for the production of sustainable fuel. ChemSusChem 2008, 1, 179.
(108) Zhu, X.; Lobban, L. L.; Resasco, D. E.; Mallinson, R. G. Effects
(129) Graça, I.; Ramôa Ribeiro, F.; Cerqueira, H. S.; Lam, Y. L.;
of HZSM-5 crystallite size on stability and alkyl-aromatics product
Almeida, M. B. B. Catalytic cracking of mixtures of model bio-oil
distribution from conversion of propanal. Catal. Commun. 2010, 11,
compounds and gasoil. Appl. Catal. B: Environ. 2009, 90, 556.
977.
(130) Fogassy, G.; Thegarid, N.; Toussaint, G.; van Veen, A. C.;
(109) Chantal, P.; Kaliaguine, S.; Grandmaison, J. L.; Mahay, A.
Schurmann, Y.; Mirodatos, C. Biomass derived feedstock co-
Production of hydrocarbons from aspen poplar pyrolytic oils over
processing with vacuum gas oil for second-generation fuel production
HZSM-5. Appl. Catal. 1984, 10, 317.
in FCC units. Appl. Catal. B: Environ. 2010, 96, 476.
(110) Sharma, R. K.; Bakhshi, N. N. Catalytic upgrading of pyrolysis
(131) Fogassy, G.; Thegarid, N.; Schurmann, Y.; Mirodatos, C. From
oil. Energy Fuels 1993, 7, 306. biomass to bio-gasoline by FCC co-processing: Effect of feed
(111) Adjaye, J. D.; Bakhshi, N. N. Upgrading of a wood-derived oil composition and catalyst structure on product quality. Energy Environ.
over various catalysts. Biomass Bioenergy 1994, 7, 201. Sci. 2011, 4, 5068.
(112) Adjaye, J. D.; Bakhshi, N. N. Production of hydrocarbons by
catalytic upgrading of a fast pyrolysis bio-oil. Part I: Conversion over
various catalysts. Fuel Process. Technol. 1995, 45, 161.

287 dx.doi.org/10.1021/ie301714x | Ind. Eng. Chem. Res. 2013, 52, 275−287

Das könnte Ihnen auch gefallen