Вы находитесь на странице: 1из 9

Bioresource Technology xxx (2015) xxxxxx

Contents lists available at ScienceDirect

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

Review

Various pretreatments of lignocellulosics


Harifara Rabemanolontsoa, Shiro Saka
Department of Socio-Environmental Energy Science, Graduate School of Energy Science, Kyoto University, Yoshida-honmachi, Sakyo-ku, Kyoto 606-8501, Japan

h i g h l i g h t s
 This review focuses on the development of pretreatment technologies for biomass.
 Achievements and current approaches on chemical, physico-chemical and biological pretreatments are reported.
 Advantages and drawbacks of the existing pretreatment technologies are discussed.
 Novel single-step and effective pretreatment methods with a common solvent and/or performant microorganisms are desired.

a r t i c l e

i n f o

Article history:
Received 6 July 2015
Received in revised form 9 August 2015
Accepted 10 August 2015
Available online xxxx
Keywords:
Biomass
Pretreatment
Chemical
Physico-chemical
Biological

a b s t r a c t
Biomass pretreatment for depolymerizing lignocellulosics to fermentable sugars has been studied for
nearly 200 years. Researches have aimed at high sugar production with minimal degradation to inhibitory compounds. Chemical, physico-chemical and biochemical conversions are the most promising technologies. This article reviews the advances and current trends in the pretreatment of lignocellulosics for a
prosperous biorefinery.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
The alarming environmental, economic and social issues engendered by massive use of fossil resources have encouraged intensive
researches on substitute raw materials for energy, materials and
chemical production. Alternative energy production can be implemented by the use of different renewables such as wind, water and
sun, but the industries based on sustainable materials, chemicals
and fuels rely mostly on lignocellulosic biomass.
Lignocellulosics are abundantly available, relatively distributed
worldwide and may alleviate the conflict in use between food and
energy. Bioconversion of lignocellulosics to liquid and gases is one
of the prospective approaches for sustainable biofuels, biochemical
and biomaterials as combined in a concept called biorefinery. The
traditional microorganisms used to produce the targeted products
such as ethanol cannot directly ferment the polymeric lignocellulosics. Thus, a pretreatment is necessary and essential to hydrolyze
the lignocellulosics into fermentable sugars.

Corresponding author. Tel./fax: +81 75 753 4738.


E-mail address: saka@energy.kyoto-u.ac.jp (S. Saka).

Pretreatment technologies for lignocellulosics have their origin


in wood science for pulp and paper fabrication but also in agricultural and crop science to increase the digestibility of forage by
ruminants. They can be retraced from 1819 when the French plant
chemist H. Braconnot discovered that glucose is formed by treating
wood with concentrated sulfuric acid (Braconnot, 1819). Since
then, various pretreatment technologies have been proposed, challenging the complexity of biomass structure and attempting to
overcome the duality of obtaining high recovery of fermentable
products with no or very limited amount of degradation products.
In the last decades, numerous pretreatment researches have
been accomplished, focusing on the identification, evaluation and
demonstration of promising approaches. The selection and the
chemistry of pretreatments affect the biorefinery system in terms
of configuration and cost of the global process. To ease the choice
of a pretreatment for a given application, the present work synthesizes the technical evolution and recent developments on the most
promising pretreatment technologies, including chemical, physicochemical, and biological procedures. The factors affecting the treatments are discussed. Furthermore, these different technologies are
assessed in their advantages and disadvantages as well as their
effects on the treated biomass.

http://dx.doi.org/10.1016/j.biortech.2015.08.029
0960-8524/ 2015 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

2. Chemical and physico-chemical pretreatments

2.2. Alkaline pretreatment

Most reviews separate chemical and physico-chemical pretreatments, while in fact, the physico-chemical processes are improvements of the established chemical processes, to diminish reaction
time and improve efficiency. Therefore, for a better understanding
of the technical evolution, they are synthesized under the same
section in this work.

2.2.1. Chemical alkaline pretreatment


Alkaline pretreatment of lignocellulosics originates from soda
pulping patented in 1854 (Watt, 1854). In biorefinery, it is essentially used for delignification to enhance the accessibility and
digestibility of polysaccharides prior to enzymatic hydrolysis or
consumption by microorganisms. The main alkaline reagents are
sodium hydroxide (NaOH), potassium hydroxide (KOH), aqueous
ammonia (NH4OH), calcium hydroxide (Ca(OH)2) and oxidative
alkali.
Compared to other chemical pretreatment technologies, alkaline hydrolysis can be conducted at lower temperature and pressure causing less sugar degradation than acid pretreatment but
the reaction times take several hours or days, or even weeks for
softwood (Bali et al., 2015).
As a mechanism, the lignocellulosic cell wall is swollen and the
internal surface area increases. In the meantime, the alkali agent is
believed to saponify the uronic ester linkages of 4-O-methyl-

2.1. Acid pretreatment


In this chemical pretreatment method, acids are used as catalysts. At first, inorganic acids such as sulfuric, hydrochloric,
hydrofluoric, phosphoric, nitric acids etc. were applied and the oldest one reported in 1819 is concentrated sulfuric acid for saccharification of cellulose (Braconnot, 1819). After such an initial
discovery, various concentrated or dilute inorganic and organic
acids were assessed to hydrolyze lignocellulosics.
Concentrated acid allows to obtain high yield of sugars such as
glucose from cellulose with low temperature. Hydrolysis rate is,
however, slower for crystalline cellulose than for amorphous hemicellulose due to their difference in intrinsic properties. Consequently, when hydrolysis is performed in one step,
hemicellulose-derived pentoses and hexoses are more susceptible
for decomposition to furfural and 5-hydroxymethyl furfural
(5-HMF) etc. which are known to have inhibitory effects on the
subsequent fermentation of the sugars. Other disadvantages of
concentrated acid include corrosion of the equipment, high consumption of the acid, toxicity to the environment, and energy
demand for acid recovery (Jones and Semrau, 1984).
As compared to concentrated acid, dilute acid hydrolysis presents the advantage of lower acid consumption but in return,
higher temperature is required and strong conditions should be
applied to achieve reasonable yield of glucose from crystalline cellulose, resulting in an extensive degradation of the amorphous
hemicelluloses.
In an effort to overcome the drawbacks of concentrated and
diluted acids, a 2-stage processes has been developed using sulfuric acid. In the first stage, 70% sulfuric acid at 3040 C is used to
hydrolyze hemicellulose and decrystallize cellulose. The acid is
then diluted to be 3040% with hot water for the second stage,
and the temperature is increased to 9095 C to hydrolyze the
decrystallized cellulose. At the end of the hydrolysis, an ion
exchange column is used to separate sulfuric acid from the
hydrolyzates obtained after the 2 stages, and the isolated sulfuric
acid is recycled for further acid hydrolysis. This concentrated sulfuric acid process is now under way for commercial application
(Saka, 2009).
Recently, organic acids were suggested as alternatives to inorganic ones in order to avoid machine corrosions and to lower
energy demand for acid recovery. Maleic, succinic, oxalic, fumaric
acids as dicarboxylic acids and acetic acid as monocarboxylic one
have been evaluated. Even at similar pH values, monocarboxylic
acid was found to have lower catalytic performance than dicarboxylic acids due to their difference in pKa (Trzcinski and
Stuckey, 2015). Maleic acid can hydrolyze cellulose as effectively
as sulfuric acid and it does not promote degradation reactions
because its double pKa values favor cellulose hydrolysis over glucose decomposition. Consequently, glucose decomposition is much
lower during cellulose hydrolysis with maleic acid in comparison
to sulfuric acid but hemicellulosic sugars are more extensively
decomposed (Lee and Jeffries, 2011). Organic acids would, thus,
be better for biomass with high cellulose and lower hemicellulose
contents such as aquatic plants (Rabemanolontsoa and Saka, 2012).

D-glucuronic acids attached to the xylan backbone, producing a


charged carboxyl group and cleaving the linkages to lignin and
other hemicelluloses. The resulting removal of lignin and hemicellulose destroys the cellulosehemicelluloselignin matrix and provokes the disruption of hydrogen bonds in cellulose (McMillan,
1994). Through scanning electron microscope, X-ray diffraction
and Fourier transform infrared spectrometer observations, biomass
pretreated with NaOH aqueous solution showed an increase in
porosity and a greater surface area (Janu et al., 2011). Bali et al.
demonstrated recently through comparison of various pretreatment methods that the highest increase in cellulose accessibility
was with dilute NaOH solution, followed by methods using NH4OH
soaked in Ca(OH)2 solution (Bali et al., 2015).
The digestibility of alkali-pretreated biomass is reported to be
inversely proportional to the lignin content (Millett et al., 1976).
Due to the difference in intrinsic properties of lignin for angiosperms and gymnosperms, alkaline hydrolysis is more effective on
hardwoods and agricultural residues in angiosperms which have
lower lignin content, than softwoods in gymnosperms with lignin
content higher than 26% (Millett et al., 1976).
In a similar way as for acid pretreatment, the alkali should be
recovered after the treatment, engendering additional energy
input. To overcome this issue, the use of ammonia gas was developed lately and 90% of the anhydrous ammonia used to pretreat
the biomass could be removed by simple ventilation. The remaining ammonia was suggested to improve ethanol production by
supplying assimilable nitrogen for microbial growth during the
subsequent fermentation (Yoo et al., 2011).
Other drawbacks of alkaline hydrolysis are the use of high temperature which increases energy demand and removes part of
hemicellulose, causing sugar loss. Under such a situation, soaking
the biomass in 15 or 30 wt% aqueous ammonia at low temperature
(3075 C) for 12 h to 77 days seems to be an interesting option
(Kim et al., 2008). The process is called soaking aqueous ammonia
(SAA). Although it was previously claimed that hemicellulose is
retained in the solid residue together with glucan, a recent study
has shown that some loss in hemicellulose and cellulose still
occurs, especially for grass (Antonopoulou et al., 2015). However,
the loss is minor as compared to other procedures and sugar degradation is minimal. The process is ideal when both hexoses and pentoses are planned to be used as substrates for the subsequent
fermentation after enzymatic hydrolysis.

2.2.2. Physico-chemical alkaline pretreatment


In order to enhance lignin removal and increase efficiency,
physical parameters such as pressure were added to the chemical

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

alkaline pretreatment. The mostly applied ones are Ammonia-fiber


expansion (AFEX) and Ammonia recycle percolation (ARP).
AFEX is a physico-chemical pretreatment conducted using
12 kg of liquid ammonia per kg of dry biomass in relatively moderate temperature (60120 C) and high pressure (1.722.06 MPa)
for 530 min, and then the pressure is suddenly reduced (Kumar
et al., 2009). It is similar to steam explosion but liquid ammonia
is used instead of water.
Dale (1986) is one of the first to patent the technology in 1986
in the direction of improving forage ammoniation which has been
developed since 1905 for lignin removal in forages. The most influencing parameters in AFEX are ammonia and water loadings, temperature, blowdown pressure, time and number of treatments
(Behera et al., 2014).
The combination of alkali agent and high pressure creates a
physico-chemical alteration in the ultra- and macro-structures of
biomass. In a similar way as for ordinary alkaline hydrolysis, there
is almost no sugar loss and the pretreatment results in decrystallization of cellulose, partial depolymerization of hemicellulose,
removal of acetyl group from hemicellulose, cleavage of lignincarbohydrate complex (LCC) linkages and the COC bonds in lignin.
The pressurization increases the surface area and the wettability
as compared to regular alkaline hydrolysis. All these effects contribute to a better accessibility for the enzymes and an improved
digestibility of the pretreated biomass resulting in a sugar recovery
over 90% after enzymatic hydrolysis of Bermuda grass (containing
approximately 5% lignin) and bagasse (15% lignin). However, AFEX
is still not ineffective on biomass with high lignin content such as
aspen chips, nutshell etc., showing sugar recovery up to 50% after
enzymatic hydrolysis (McMillan, 1994). Additionally, AFEX uses a
considerable amount of ammonium which necessitates high
energy input for recovery and recycling.
Ammonia recycle percolation (ARP) is another form of alkaline
physico-chemical treatment in which aqueous ammonia (2.5
15%) passes through a reactor packed with biomass, with a temperature set between 140210 C. The reaction time is about 90 min
and the percolation rate around 5 ml/min. Although ARP is efficient
for delignification of hardwoods and herbaceous plants, it is less
effective for softwoods (Chandra et al., 2007).
2.3. Hydrothermal pretreatment
The term hydrothermal was defined for the first time by the British geologist Sir Roderick Murchison (17921871) to describe the
action of water at elevated temperature and pressure (Yoshimura
and Byrappa, 2008). In spite of different definitions proposed by
various scientists afterwards, hydrothermal treatment defines
commonly as reactions occurring under the conditions of high
temperature and high pressure in aqueous solutions in a closed
system. In more detail, hydrothermal treatments include steam
explosion, supercritical/subcritical water and hot-compressed
water treatments depending on the conditions of temperature
and pressure.
2.3.1. Steam explosion
Steam explosion is a physico-chemical pretreatment. It is
known since the 1800s but the earliest patents can be retraced to
1924 for chipboard production from wood (Mason, 1926). The process has been used since the second half of the 20th century for the
production of feed for ruminants and nowadays it is the most commonly used method for hydrothermal pretreatment of woody biomass (McMillan, 1994).
It is a thermo-mechano-chemical treatment using highpressure saturated steam at about 0.694.83 MPa with temperature ranged between 160 and 260 C for several seconds to a few
minutes. The steam penetrates into the treated biomass and

expands the cell walls of the fibers prior to explosion and partial
hydrolysis. Then, the pressure is rapidly reduced down to atmospheric condition.
During this treatment, acetyl residues from xylan hemicellulose
are liberated in form of acetic acid and catalyze the chemical reaction. The phenomenon is qualified to be autohydrolysis. It was
reported that 90% efficiency of enzymatic hydrolysis can be
achieved after steam explosion of poplar (Populus tremuloides)
(Grous et al., 1986). Similar positive effects were reported on other
hardwood, pine chips, French maritime pine (Pinus pinaster), rice
straw, bagasse, olive stones, giant miscanthus (Miscanthus  giganteus) and spent Shiitake mushroom media (Jacquet
et al., 2012).
Temperature, residence time, chip size and moisture content
are the most important parameters affecting steam explosion,
and it was found that lower temperature and long residence time
were more favorable (Wright, 1988).
Steam explosion without catalyst was less performant on softwood and the technique was improved by addition of acid catalyst
during the treatment or soaking into acid before the treatment
(Eklund et al., 1995).
Steam explosion is one of the most cost-effective processes with
lower energy requirement as compared with mechanical milling,
but it cannot achieve complete disruption of the lignincarbohydrate matrix, and thus, products from hemicellulose are not completely recovered. Other drawbacks include the production of
inhibitory compounds to microorganisms and enzymes in the
downstream processes, so that removal of the inhibitory products
becomes necessary before enzymatic hydrolysis (McMillan, 1994).

2.3.2. Supercritical and subcritical water


Water is one of the most important solvents abundantly present
in nature. It presents remarkable safety advantages because it is
environmentally benign, non-toxic, non-flammable, noncarcinogenic, non-mutagenic, and thermodynamically stable.
The phase and chemical properties of water change depending
on the conditions of temperature and pressure, as shown in
Fig. 1. In this figure, the critical point represents water at its critical
temperature (Tc = 374 C) and pressure (Pc = 22.1 MPa). Above the
critical point, when both temperature and pressure exceed Tc and
Pc, respectively, water becomes supercritical and presents a noncondensable gas-like state with a density reduced to half until
one third of water in ambient condition.

Fig. 1. Changes in phase of pure water according to temperature and pressure.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

While the supercritical water is clearly defined, the subcritical


state of water is less precise in temperature and pressure conditions. When temperature and/or pressure do not exceed Tc and/
or Pc, water is called to be subcritical. Therefore, only in a vicinity
of the critical point of water is called as subcritical water, while
hydrothermal condition far away from critical point, thus milder
condition, is defined as hot-compressed water.
In subcritical and supercritical states, the ionic product of water
and its dielectric constant can be easily controlled by changing
temperature and pressure. In consequence, properties of water
can be continuously and largely changed, ranging from aqueous
solution to non-aqueous one. In supercritical and subcritical state
particularly, the ionic products of water increase the hydrolytic
degradation capacity and thus hydrolysis reactions can be achieved
(Saka, 2001).
With subcritical or supercritical water, the ester and ether linkages in chemical constituents of lignocellulosics can be hydrolyzed
in the absence of catalyst (Sasaki et al., 1998) and useful lowmolecular-weight chemicals can be obtained.
The decomposition pathway of cellulose as treated by flow-type
supercritical water at 380 C/40 MPa/0.120.48 s is shown in Fig. 2
(Ehara et al., 2002). At first, cellulose is hydrolyzed to polysaccharides which degree of polymerization (DP) is in a range between 13
and 100. The obtained polysaccharides are further hydrolyzed to
oligosaccharides with DP between 2 and 12. The reducing ends
of the resulting oligosaccharides are then dehydrated to levoglucosan and fragmented to erythrose/glycolaldehyde as revealed by
analyses with Matrix-Assisted Laser Desorption/Ionization-Timeof-Flight Mass Spectrometry (MALDI-TOFMS) analysis (Ehara
et al., 2002). The non-reducing ends of the oligosaccharides are
hydrolyzed to glucose, which, to some extent, can be isomerized
to fructose. If the treatment is prolonged, the resultant hexoses
may be decomposed to levoglucosan, 5-HMF, erythrose, glycolaldehyde, methylglyoxal, and dihydroxyacetone by way of dehydration or fragmentation. The dehydrated and fragmented
products can further be oxidized to low-molecular-weight organic
acids such as pyruvic, lactic, formic and acetic acids. Under high
supercritical temperature and pressure around 400650 C/28
34.5 MPa, the biomass is gasified and the main products are H2,
CO2 and CH4 (Antal et al., 2000).

The hydrolysis reaction of lignocellulosics is so fast in the order


of seconds that controlling the reaction to obtain more hexoses and
less degradation products is rather difficult. Under such a situation,
a sudden expansion micro-reactor which allows the instantaneous
cooling of the products efficiently stops the hydrolysis reactions at
very short times. Precise evaluation of the reaction time without
diluting the products is, thus, possible. When this method was
applied to wheat bran, a total recovery of hemicellulosic pentoses
was achieved at 0.19 s and 65 wt% of cellulosic hexoses were
obtained after 0.22 s (Cantero et al., 2015b).
Reaction time and ion concentration in the reaction medium
govern the selectivity of the process. In other words, dissociated
water molecules diminish the dehydration of hexoses (Cantero
et al., 2015a). As for the fate of lignin during supercritical treatment, the ether linkages are easily cleaved, whereas the condensed
type linkages of lignin are rather stable (Ehara et al., 2002).
From total ion chromatogram of the methanol-soluble portion
of Japanese cedar (Cryptomeria japonica) as treated by supercritical
water, oily substances as lignin-derived products could be collected and all these products have guaiacyl nuclei (2methoxyphenol), except for 5-HMF, a contaminant from cellulose.
The obtained phenylpropane units (C6C3, together with C6C2,
C6C1) derived from the cleavage of ether linkages of lignin are
useful alternatives to aromatic chemicals from fossil resources.
Cleavage between Cb and Cc (Cb/Cc) bond in the C6C3 unit and
dealkylation of propyl chain of alkyl phenols were also observed
during supercritical water treatments (Ehara et al., 2002).
Although flow-type supercritical water treatment provides high
yield of hydrolyzed products, part of the sugar is lost due to cellulose fragmentation (Ehara and Saka, 2002, 2005). On the other
hand, cellulose under subcritical water is likely to be less decrystallized. A combined supercritical and subcritical treatments were
therefore developed and increased the yield of saccharides with
fewer degradation products (Ehara and Saka, 2005).
2.3.3. Hot-compressed water
Hot-compressed water, also known as liquid hot water, has
been explored as a milder condition compared to sub- and supercritical water. With temperatures above 200 C at various pressure
conditions below the critical point, hot-compressed water presents

Fig. 2. Proposed pathway of cellulose decomposition as treated by a flow-type supercritical water.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

the advantage to act as a solvent and reaction medium. Various


studies on hot-compressed water hydrolysis were performed to
investigate the hydrolysis performance (Abdullah et al., 2014),
reaction mechanisms (Abdullah et al., 2013) and kinetics (Kamio
et al., 2008). Due to the difference in hydrolysis rate for cellulose
and hemicellulose, the treatment cannot be optimized under the
same severity so that two-step hot-compressed water treatment
has been proposed. The first stage is performed at low severity to
hydrolyze hemicelluloses while the second stage, at a higher severity aims to depolymerize cellulose and increase the yield of
saccharides.
Two-step hydrolyses (230 C/10 MPa/15 min and 270280
C/10 MPa/15 min) of Japanese beech (Fagus crenata), Japanese
cedar, Nipa (Nypa fruticans) frond and rice straw achieved to solubilize respectively 92.2%, 82.3%, 92.4% and 97.9% of the starting
biomass (Ogura et al., 2013; Phaiboonsilpa, 2010). The treatment
is therefore applicable to a wide range of biomass species including
softwoods.
Hydrolyzed products in the first stage are xylo-oligosaccharides,
xylose, glucuronic acid and acetic acid from O-acetyl-4-Omethylglucuronoxylan (Nakahara et al., 2014). Other monomeric
and oligomeric sugars such as glucomanno-saccharides, mannose,
rhamnose, arabinose etc. from other types of hemicellulose are also
obtained, depending on the starting biomass. For lignin, on the
other hand, hydrolyzed monomeric guaiacyl and syringyl units
and their dimeric condensed type units were recovered, while
the hydrolyzed products in the second stage were cellosaccharides such as glucose and cello-oligosaccharides from cellulose. Although part of the lignin is liquefied during the treatment,
the rest remains in the residue in condensed type structures
(Takada and Saka, 2015), facilitating its separation from the hydrolyzate and its further use in biorefinery.
Under the milder condition of hot-compressed water, the
reducing end of oligosaccharides can be dehydrated to levoglucosan, but sugar decomposition is minimized and the concentration of furfural or 5-HMF did not exceed 1.5% (Ogura et al., 2013;
Phaiboonsilpa, 2010).
2.4. Ionic liquid
Ionic liquids are new organic salts which generally exist in liquid state at ambient temperature due to their low melting point.
Intrinsically useful characteristics such as high ionic conductivity,
high solvation power, thermal stability, inflammability, low
volatility and recyclability confer the status of green solvent on
ionic acid for various chemical reactions in industrial processes.
An asymmetric organic cation with an organic or inorganic
counterpart usually composes an ionic liquid. Since a wide range
of anions and cations can be conceived to design a particular ionic
liquid, the properties can be tailored. Examples of cations combined
with inorganic anions include 1-ethyl-3-methylimidazolium chloride, 1-butyl-3-methylimidazolium chloride and 1-allyl-3methylimidazolium chloride. An example of cation with organic
anion is 1-ethyl-3-methylimidazolium acetate which was used to
liquefy wood (Clough, 2015).
Efforts to clarify the detailed behaviors and nature of reactions
during the liquefaction of wood in ionic liquids were accomplished
recently. A study on topochemical and morphological characterization of wood cell wall treated with 1-ethylpyridinium bromide
showed that the ionic liquid is more reactive toward lignin than
polysaccharides, resulting in inhomogeneous changes in ultrastructural and chemical composition of the cell wall (Kanbayashi
and Miyafuji, 2015). Moreover, thermodynamic studies indicated
that liquefaction of cellulose in imidazolium acetate ionic liquid
is exothermal but an increase in temperature does not thermodynamically favor the liquefaction process (Andanson et al., 2015).

The addition of water (above 35 wt%) or solvent such as dimethyl


sulfoxide (DMSO) renders the reaction endothermic (de Oliveira
and Rinaldi, 2015). It was also demonstrated that cation plays a
secondary role compared with the anion in the overall energetics
of the system (de Oliveira and Rinaldi, 2015).
Despite their promising chemical properties, ionic liquids present the drawback of being expensive and require tedious recycling, since their toxicity and biodegradability are not yet well
understood. Therefore, further research on these aspects are
necessary.

3. Biological pretreatments
3.1. Biological saccharification
3.1.1. Microbial saccharification
A great diversity of bacteria and fungi can hydrolyze cellulose
and hemicellulose into their monosugar counterparts. While cellulose hydrolyzing species are widely distributed among fungi, this
capacity in bacteria is mostly found in Clostridiales (anaerobic)
and Actinomycetales (aerobic). Table 1 shows the main hydrolytic
microorganisms with their culture conditions and performances.
Thermophilic and extremophilic microorganisms, with optimum
temperature above 50 C are preferred for biorefinery because they
show better adaptation to pH, temperature and environmental
changes. Also, processing at higher temperature reduces the development of eventual mesophilic contaminants.
Among the hydrolytic microorganisms, clostridia are the most
investigated. Under anaerobic condition, they grow on the surface
of cellulosic materials and degrade the polysaccharides by means
of an extra-cellular complex enzyme system called cellulosome.
The variety of enzymes associated with cellulosome presently
known includes endoglucanases, exoglucanases, hemicellulases,
chitinases, pectin lyases, and lichenases (Bayer et al., 2004).
In the past few decades, cellulolytic clostridia have been intensively studied for lignocellulosic biofuel production. Among them,
the thermophilic Clostridium thermocellum shows the highest
growth rate on crystalline cellulose (Lynd et al., 2002).
In contrast, aerobic cellulolytic fungi and bacteria do not have
cellulosome. Instead, they produce free cellulases, which by hyphal
extensions, can penetrate the lignocellulosic substrates and hydrolyze them (Chang and Yao, 2011). While the spatial arrangements
of enzymes in anaerobic and aerobic microorganisms can be different, they use the same group of enzymes and their rate of hydrolysis is similar (Dionisi et al., 2015; Lynd et al., 2002).
The main important parameters for microbial saccharification
are temperature, pH, particle size, substrate accessibility and
hydrogen partial pressure (Dionisi et al., 2015).
As can be seen in Table 1, for both anaerobic and aerobic
microorganisms, the optimum pH values range from 6.5 to 8 and
cellulolytic activities usually stop at pH bellow 5.5. Faster hydrolysis rate can be achieved with smaller particle size (Hu et al.,
2005). Hydrogen concentration in the liquid phase and hydrogen
partial pressure in the headspace were demonstrated to affect
the spectrum of product ratio, particularly in anaerobic condition
(Dionisi et al., 2015). High hydrogen concentration can lead to
the formation of more acids, alcohols, CO2 and H2 (Lamed et al.,
1988), instead of monosaccharides.
Indeed, monosaccharides are intermediate compounds in the
metabolism of cellulolytic microorganisms and it is usually difficult, though not impossible (Prawitwong et al., 2013), to stop the
metabolism to the monosaccharides level. Therefore, enzymatic
saccharification has been explored, where only the hydrolytic
enzymes extracted from the microorganisms would be used in
order to easily obtain monosaccharides only as end-products.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

Table 1
Culture conditions and performances of various cellulolytic, hemicellulolytic and/or ligninolytic microorganisms.
Microorganisms

Activity pH

Temperature Treatment time Degradation References


(C)
(days)
(%)

C, H

6.17.8

60

45

85100

Rabemanolontsoa et al. (2015)

C, H
C, H

6.07.5
7.0

95
8085

C, H

7.2

3234

36

2075

Wu et al. (2011) as reviewed by Khare et al. (2015)


Huser et al. (1986), Ruttersmith and Daniel (1991) as reviewed by
Khare et al. (2015)
Desvaux et al. (2001) as reviewed by Dionisi et al. (2015)

C, H

6.0

37

Koukiekolo et al. (2005)

C, H
C, H

6.77.1
6.56.8

37
39

0.52
27

3070
5487

C, H

6.16.84

38

0.53

5479

Pavlostathis et al. (1988) as reviewed by Dionisi et al. (2015)


Pettipher and Latham (1979), Shi and Weimer (1992) as reviewed
by Dionisi et al. (2015)
Roger et al. (1990) as reviewed by Dionisi et al. (2015)

C, H

6.5

3055

28

60

Bagnara et al. (1987)

C, H
C, H

2.57
34

8890
60

Zillig et al. (1981) as reviewed by Khare et al. (2015)


Eckert et al. (2002) as reviewed by Khare et al. (2015)

C, H
C, H

7.07.2
6.6

30
35

4
35

100
1570

C, H
C, H
L

5.0
4.8
5.37.8

30
28
30

0.51.2
7
760

5075
100
2052

Xanthomonas spp.
L
Acinetobacter spp.
L
Streptomyces cyaneus C, H, L

30
30
2837

730
30
2128

3948
4757
2952

Thermomonospora
mesophila
Pleurotus ostreatus

37

21

3648

Li and Gao (1997) as reviewed by Dionisi et al. (2015)


de Coninck-Chosson (1988), Rapp and Wagner (1986) as reviewed
by Dionisi et al. (2015)
Peitersen (1977) as reviewed by Dionisi et al. (2015)
Velkovska et al. (1997) as reviewed by Dionisi et al. (2015)
Srensen (1962), Vidal et al. (1989) as reviewed by Dionisi et al.
(2015)
Odier et al. (1981) as reviewed by Dionisi et al. (2015)
Odier et al. (1981) as reviewed by Dionisi et al. (2015)
Berrocal et al. (2000), Zimmermann and Broda (1989) as reviewed
by Dionisi et al. (2015)
Zimmermann and Broda (1989) as reviewed by Dionisi et al. (2015)

2530

3060

4041

Phanerochaete
L
chrysosporium
Echinodontium taxodii L
2538
Trametes versicolor
L
spp.

39

1430

2860

25

28

24

Kerem et al. (1992), Taniguchi et al. (2005) as reviewed by Dionisi


et al. (2015)
Kerem et al. (1992), Shi et al. (2008) as reviewed by Dionisi et al.
(2015)
Zhang et al. (2007) as reviewed by Dionisi et al. (2015)

25

28

924

Zhang et al. (2007) as reviewed by Dionisi et al. (2015)

Anaerobic
Clostridium
thermocellum
Thermotoga maritima
Thermotoga strain
FjSS3-B.1
Clostridium
cellulolyticum
Clostridium
cellulovorans
Ruminococcus albus
Ruminococcus
flavefaciens
Fibrobacter
succinogenes
Facultative anaerobic
Actinotalea
fermentans
Aerobic
Sulfolobus solfataricus
Alicyclobacillus
acidocaldarius
Cytophaga sp. LX-7
Cellulomonas uda JC3
Trichoderma viride
Trichoderma reesei
Pseudomonas spp.

C: cellulase, H: hemicellulase, L: ligninase.

Besides, isolating the enzymes enables to produce multiple


enzyme mixtures and the best combination would be used to
obtain the targeted products in the shortest time.
3.1.2. Enzymatic saccharification
The main microbial enzyme sources are shown in Table 1. For
cellulose hydrolysis, cellulases are the primary enzymes. They
are composed of three predominant activites: exo-1,4-bglucanase (EC 3.2.1.91), endo-1, 4-b-glucanase (EC 3.2.1.4), and cellobiase also known as b-glucosidase (EC 3.2.1.21). The efficiency of
saccharification depends on the balanced and appropriate combination of these activities as well as the type of biomass to be pretreated (Khare et al., 2015).
For hemicellulose hydrolysis, xylanases are the key enzymes
and the mechanism known so far is the following. First, endo-b1,4-xylanase depolymerizes the xylan backbone and the resulting
xylooligosaccharides are further degraded by b-xylosidase, while
side groups are cleaved by different glycosidases such as
b-glucuronidases, a-arabinofuranosidase, acetyl xylan esterase,
and ferulic acid esterase (Saha et al., 2005). Some synergistic interactions between endoxylanases and glucosidases and between glucuronidase and xylosidase (Kormelink and Voragen, 1993;

Srensen et al., 2003) indicated that (1) the removal of side groups
enhances the activity of endoxylanases and (2) the unsubstituted
xylooligomers (products of glucuronidase) are better substrates
for xylosidase. Despite successful isolation and characterization
of different enzymes, hemicellulase systems are less understood
compared with the cellulase systems. Further studies on hemicellulose degradation and its regulation are highly expected in the
future to improve the lignocelluloses degradation.
3.2. Biological delignification
3.2.1. Microbial delignification
One of the earliest conclusive discoveries on lignin biodegradation can be retraced from 1939 on compost studies by Waksman
and Cordon (1939). Further evidences published in 1957 supported
the biological removal of lignin from lignocellulosic matrix by
white-rot fungi (Lawson and Still, 1957). Then, numerous studies
on delignification by different fungal and bacterial species followed
such pioneering findings and are synthesized in different reviews.
Plcido et al. published a recent one (Plcido and Capareda, 2015).
The pretreatment technology involving delignification of lignocellulosics with microorganisms is called solid-state fermentation.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

The most studied ligninolytic species are white-rot fungi, with


their enzymatic system composed of laccase, manganese peroxidase (MnP) and lignin peroxidase (LiP). The rate with which fungi
attack lignin and carbohydrates in lignocellulosic tissues greatly
depends on the strain used as well as on the amount and nature
of lignin and polysaccharides in the substrate.
White-rot fungi preferentially degrade syringyl (S) units of lignin, while guaiacyl (G) units are more resistant (Hatakka, 2005).
For example, the fungus Phlebia radiata removed lignin from the
secondary wall of wheat straw, consisting mainly of S lignin
(Burlat et al., 1998). Even with the extensive researches on the fungal degradation of lignin, the catabolic pathways are still incomplete, and in practice are hard to elucidate, due to the complex
structure of lignin, and its resistance to hydrolysis. Furthermore,
delignification efficiencies of lignocellulosics with fungi range only
from 15% to 39.2%, while treatment times reach 146 days (Plcido
and Capareda, 2015).
Apart from fungi, certain bacteria for instance, Pseudomonas
and filamentous bacteria known as Actinomycetes, have the capacity to delignify plant cell wall with the help of their enzymes such
as peroxidase, MnP, LiP, demethylase, catalase and phenol oxidase
(Hatakka, 2005). Typical mechanisms include solubilization of lignin and production of acid-precipitable polymeric lignin (APPL),
enzymatic cleavage of CaCb bonds by radicals generated from oxidation of Ca in the side chain of carbonyl group, alkyl-aryl cleavage,
cross-linking, demethylation and formation of monomeric products such as 4-ethoxy-3-methoxybenzaldehyde, guaiacol, vanillic
acid and protocatechuic acid (Godden et al., 1992). Certain bacteria
can even achieve ring cleavage of aromatic compounds (Salvachua
et al., 2015).
Although external factors such as temperature, pH and moisture content are important parameters, microbiological delignification depends mainly on the strain of microorganism used, their
selectivity and their substrate specialization (hardwood, softwood
or herbaceous species). For example, Ganoderma lucidum significantly produced MnP on poplar wood but not on pine wood under
the same conditions (DSouza et al., 1999). Applying fungal delignification prior to acid hydrolysis seems to be judicious since some
organic acids secreted by the fungi reduce the pH and consequently diminish the acid loading for hydrolysis. Less degradation
products were obtained from this process combination (Kuhar
et al., 2008).
In addition to the lignocellulosic enzyme complex, most lignocellulolytic microorganisms produce other enzymes, such as pectinases, proteases, lipases and phytases on lignocellulosic substrates
(Tengerdy and Szakacs, 2003). Therefore, together with lignin
removal, a portion of hemicellulose and cellulose is also biodegraded, engendering loss in fermentable sugars. Selective microorganisms are few, and they are all white-rot fungi (Hatakka, 2005).
Isolation of the ligninolytic enzymes and their separate utilization
is another solution to improve selectivity.
3.2.2. Enzymatic delignification
Yoshida (1883) was the first to isolate ligninolytic enzymes in
1883 when he extracted laccase from the Japanese lacquer-tree
(Rhus vernicifera). Then, about a hundred years later, Kuhara separated MnP from batch cultures of Phanerochaete chrysosporium
(Kuwahara et al., 1984). Up to now, 4 ligninolytic enzymes have
been successfully isolated: laccase, MnP, LiP and versatile peroxidase (Pollegioni et al., 2015).
Laccase was found in different kinds of plants such as Japanese
lacquer tree, mango, mung bean, peach, tobacco, maize etc. where
they participate in the xylem tissue lignification. Ligninolytic
enzymes can also be extracted from fungi, bacteria [195, 215
217] and insects such as leafhopper, hornworm and mosquito, or
termite (Plcido and Capareda, 2015). Fungal laccases are the most

widely used ligninolytic enzymes in biotechnology but pretreatment with a single enzyme cannot completely remove lignin in
either wood or herbaceous species. Laccase could remove up to
50% of lignin from wood pulp in 215 h (Annunziatini et al.,
2005), 58% from Eucalyptus globulus in 224 h (Gutirrez et al.,
2012), and 36% from herbaceous plants such as elephant grass in
24 h (Gutirrez et al., 2012). Another drawback of enzymatic pretreatment is the pH-dependence of the activity, which is optimal
at low pH, while alkaline pH would be preferred for industrial
applications (Pollegioni et al., 2015). Gene modification or use of
multiple enzymes seems to be a promising solution (Alcalde,
2015).
3.3. Consolidated bioprocessing
In recent years, the concept of consolidate bioprocessing has
emerged. It involves depolymerization of the lignocellulosic matrix
with simultaneous production of enzymes and useful products
such as ethanol or acids in one single step.
Brethauer et al. successfully achieved 67% ethanol yield from
pretreated wheat straw (dilute acid) using 3 naturally occurring
strains: Trichoderma reesei, Saccharomyces cerevisiae and Scheffersomyces stipites (Brethauer and Studer, 2014). However, the reactor
configuration was quite complex in that study and a preliminary
acid hydrolysis was necessary.
Bacterial screening for consolidated bioprocessing of lignin has
been accomplished very recently (Salvachua et al., 2015). Amycolatopsis sp., Acinetobacter ADP1 and Rhodococcus jostii were found
to be able to extracellularly depolymerize high molecular weight
lignin and intracellularly catabolize a significant portion of the
low molecular weight aromatics into fatty acids, polyhydroxyalkanoates etc., which can be used as hydrocarbon fuels or materials
precursors in biorefinery.
The use of multiple native microorganisms for consolidated bioprocessing is challenging because they do not necessarily have the
same optimum growth conditions. Another option is to engineer a
unique microorganism which can perform hydrolysis and fermentation in the same time. As summarized by den Haan et al. (2015),
researches are actively conducted to genetically engineer (i) naturally cellulolytic and/or ligninolytic microorganisms for improved
product-related properties or (ii) non-cellulolytic and/or ligninolytic microorganisms showing high product yields to express
cellulase and/or ligninase activities. Consolidated bioprocessing
as a one-pot process is still under intense investigation from laboratory to industrial scales, but it presents high economic and
technical prospective.
Overall, biological pretreatments are less harmful to the environment and can be performed at milder conditions, and thus
are energy efficient as compared with chemical and physicochemical pretreatments. They also present fewer side reactions
and necessitate less reactor resistance to pressure. Notwithstanding these advantages, improvements in the process duration, cost
reduction, increased tolerance to substrates and products are still
necessary for efficient industrial application (Plcido and
Capareda, 2015).
4. Concluding remarks
To date, a novel single pretreatment method with a common
solvent has not been established yet for delignification without
sugar degradation. The tendency is thus to use less or no catalyst
and apply successive pretreatment conditions. Interestingly, several microorganisms can directly ferment polymeric lignin, cellulose and hemicellulose into useful products. Thus, in principle, it
is possible to directly convert untreated biomass to valuable

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx

products such as ethanol, acids etc. Selection, keen combination


and optimization of these microorganisms and/or their enzymes
in a consolidated bioprocessing could be a promising approach
for successful biomass conversion.

References
Abdullah, R., Ueda, K., Saka, S., 2013. Decomposition behaviors of various crystalline
celluloses as treated by semi-flow hot-compressed water. Cellulose 20 (5),
23212333.
Abdullah, R., Ueda, K., Saka, S., 2014. Hydrothermal decomposition of various
crystalline celluloses as treated by semi-flow hot-compressed water. J. Wood
Sci. 60 (4), 278286.
Alcalde, M., 2015. Engineering the ligninolytic enzyme consortium. Trends
Biotechnol. 33 (3), 155162.
Andanson, J.M., Padua, A.A.H., Costa Gomes, M.F., 2015. Thermodynamics of
cellulose dissolution in an imidazolium acetate ionic liquid. Chem. Commun.
51 (21), 44854487.
Annunziatini, C., Baiocco, P., Gerini, M.F., Lanzalunga, O., Sjgren, B., 2005. Aryl
substituted N-hydroxyphthalimides as mediators in the laccase-catalysed
oxidation of lignin model compounds and delignification of wood pulp. J. Mol.
Catal. B Enzym. 32 (3), 8996.
Antal, M.J., Allen, S.G., Schulman, D., Xu, X., Divilio, R.J., 2000. Biomass gasification in
supercritical water. Ind. Eng. Chem. Res. 39 (11), 40404053.
Antonopoulou, G., Gavala, H., Skiadas, I., Lyberatos, G., 2015. The effect of aqueous
ammonia soaking pretreatment on methane generation using different
lignocellulosic biomasses. Waste Biomass Valorization, 111.
Bagnara, C., Gaudin, C., Belaich, J., 1987. Physiological properties of Cellulomonas
fermentans, a mesophilic cellulolytic bacterium. Appl. Microbiol. Biotechnol. 26
(2), 170176.
Bali, G., Meng, X., Deneff, J.I., Sun, Q., Ragauskas, A.J., 2015. The effect of alkaline
pretreatment methods on cellulose structure and accessibility. ChemSusChem 8
(2), 275279.
Bayer, E.A., Belaich, J.P., Shoham, Y., Lamed, R., 2004. The cellulosomes:
multienzyme machines for degradation of plant cell wall polysaccharides.
Annu. Rev. Microbiol. 58, 521554.
Behera, S., Arora, R., Nandhagopal, N., Kumar, S., 2014. Importance of chemical
pretreatment for bioconversion of lignocellulosic biomass. Renew. Sustain.
Energy Rev. 36, 91106.
Braconnot, H., 1819. Hydrolysis of cellulose into sugar. Gilberts. Ann. Phys. 63, 348.
Brethauer, S., Studer, M.H., 2014. Consolidated bioprocessing of lignocellulose by a
microbial consortium. Energy Environ. Sci. 7 (4), 14461453.
Burlat, V., Ruel, K., Martinez, A.T., Camarero, S., Hatakka, A., Vares, T., Joseleau, J.P.,
1998. The nature of lignin and its distribution in wheat straw affect the patterns
of degradation by filamentous fungi. In: Proceedings of the Seventh
International Conference on Biotechnology in the Pulp and Paper Industry,
Vancouver, BC, Canada, pp. A75A78.
Cantero, D.A., Bermejo, M.D., Cocero, M.J., 2015a. Governing chemistry of cellulose
hydrolysis in supercritical water. ChemSusChem 8 (6), 10261033.
Cantero, D.A., Martinez, C., Bermejo, M.D., Cocero, M.J., 2015b. Simultaneous and
selective recovery of cellulose and hemicellulose fractions from wheat bran by
supercritical water hydrolysis. Green Chem. 17 (1), 610618.
Chandra, R.P., Bura, R., Mabee, W.E., Berlin, A., Pan, X., Saddler, J.N., 2007. Substrate
pretreatment: the key to effective enzymatic hydrolysis of lignocellulosics? In:
Olsson, L. (Ed.), Biofuels, vol. 108. Springer, Berlin, Heidelberg, pp. 6793.
Chang, T., Yao, S., 2011. Thermophilic, lignocellulolytic bacteria for ethanol
production: current state and perspectives. Appl. Microbiol. Biotechnol. 92
(1), 1327.
Clough, M.T., 2015. Ionic liquids: not always innocent solvents for cellulose. Green
Chem.: Int. J. Green Chem. Resour.: GC 17 (1), 231243.
DSouza, T.M., Merritt, C.S., Reddy, C.A., 1999. Lignin-modifying enzymes of the
white rot basidiomycete Ganoderma lucidum. Appl. Environ. Microbiol. 65 (12),
53075313.
Dale, B.E., 1986. Method for increasing the reactivity and digestibility of cellulose
with ammonia. US Patent 4,600,590 A.
de Oliveira, H.F.N., Rinaldi, R., 2015. Understanding cellulose dissolution: energetics
of interactions of ionic liquids and cellobiose revealed by solution
microcalorimetry. ChemSusChem 8 (9), 15771584.
den Haan, R., van Rensburg, E., Rose, S.H., Grgens, J.F., van Zyl, W.H., 2015. Progress
and challenges in the engineering of non-cellulolytic microorganisms for
consolidated bioprocessing. Curr. Opin. Biotechnol. 33, 3238.
Dionisi, D., Anderson, J.A., Aulenta, F., McCue, A., Paton, G., 2015. The potential of
microbial processes for lignocellulosic biomass conversion to ethanol: a review.
J. Chem. Technol. Biotechnol. 90 (3), 366383.
Ehara, K., Saka, S., 2002. A comparative study on chemical conversion of cellulose
between the batch-type and flow-type systems in supercritical water. Cellulose
9 (34), 301311.
Ehara, K., Saka, S., 2005. Decomposition behavior of cellulose in supercritical water,
subcritical water, and their combined treatments. J. Wood Sci. 51 (2), 148153.

Ehara, K., Saka, S., Kawamoto, H., 2002. Characterization of the lignin-derived
products from wood as treated in supercritical water. J. Wood Sci. 48 (4), 320
325.
Eklund, R., Galbe, M., Zacchi, G., 1995. The influence of SO2 and H2SO4 impregnation
of willow prior to steam pretreatment. Bioresour. Technol. 52 (3), 225229.
Godden, B., Ball, A.S., Helvenstein, P., Mccarthy, A.J., Penninckx, M.J., 1992. Towards
elucidation of the lignin degradation pathway in actinomycetes. Microbiology
138 (11), 24412448.
Grous, W.R., Converse, A.O., Grethlein, H.E., 1986. Effect of steam explosion
pretreatment on pore size and enzymatic hydrolysis of poplar. Enzyme
Microb. Technol. 8 (5), 274280.
Gutirrez, A., Rencoret, J., Cadena, E.M., Rico, A., Barth, D., del Ro, J.C., Martnez, .T.,
2012. Demonstration of laccase-based removal of lignin from wood and nonwood plant feedstocks. Bioresour. Technol. 119, 114122.
Hatakka, A., 2005. Biodegradation of lignin. In: Biopolymers Online. Wiley-VCH
Verlag GmbH & Co. KGaA.
Hu, Z.-H., Yu, H.-Q., Zhu, R.-F., 2005. Influence of particle size and pH on anaerobic
degradation of cellulose by ruminal microbes. Int. Biodeterior. Biodegradation
55 (3), 233238.
Jacquet, N., Vanderghem, C., Danthine, S., Quivy, N., Blecker, C., Devaux, J., Paquot,
M., 2012. Influence of steam explosion on physicochemical properties and
hydrolysis rate of pure cellulose fibers. Bioresour. Technol. 121, 221227.
Janu, K.U., Sindhu, R., Binod, P., Kuttiraja, M., Sukumaran, R.K., Pandey, A., 2011.
Studies on physicochemical changes during alkali pretreatment and
optimization of hydrolysis conditions to improve sugar yield from bagasse. J.
Sci. Ind. Res. 70 (11), 952958.
Jones, J.L., Semrau, K.T., 1984. Wood hydrolysis for ethanol production previous
experience and the economics of selected processes. Biomass 5 (2), 109135.
Kamio, E., Sato, H., Takahashi, S., Noda, H., Fukuhara, C., Okamura, T., 2008.
Liquefaction kinetics of cellulose treated by hot compressed water under
variable temperature conditions. J. Mater. Sci. 43 (7), 21792188.
Kanbayashi, T., Miyafuji, H., 2015. Topochemical and morphological
characterization of wood cell wall treated with the ionic liquid, 1ethylpyridinium bromide. Planta, 110.
Khare, S.K., Pandey, A., Larroche, C., 2015. Current perspectives in enzymatic
saccharification of lignocellulosic biomass. Biochem. Eng. J. 102, 3844. http://
dx.doi.org/10.1016/j.bej.2015.02.033.
Kim, T.H., Taylor, F., Hicks, K.B., 2008. Bioethanol production from barley hull using
SAA (soaking in aqueous ammonia) pretreatment. Bioresour. Technol. 99 (13),
56945702.
Kormelink, F.J.M., Voragen, A.G.J., 1993. Degradation of different [(glucurono)
arabino]xylans by a combination of purified xylan-degrading enzymes. Appl.
Microbiol. Biotechnol. 38 (5), 688695.
Koukiekolo, R., Cho, H.-Y., Kosugi, A., Inui, M., Yukawa, H., Doi, R.H., 2005.
Degradation of corn fiber by Clostridium cellulovorans cellulases and
hemicellulases and contribution of scaffolding protein cbpa. Appl. Environ.
Microbiol. 71 (7), 35043511.
Kuhar, S., Nair, L.M., Kuhad, R.C., 2008. Pretreatment of lignocellulosic material with
fungi capable of higher lignin degradation and lower carbohydrate degradation
improves substrate acid hydrolysis and the eventual conversion to ethanol. Can.
J. Microbiol. 54 (4), 305313.
Kumar, P., Barrett, D.M., Delwiche, M.J., Stroeve, P., 2009. Methods for pretreatment
of lignocellulosic biomass for efficient hydrolysis and biofuel production. Ind.
Eng. Chem. Res. 48 (8), 37133729.
Kuwahara, M., Glenn, J.K., Morgan, M.A., Gold, M.H., 1984. Separation and
characterization of two extracelluar H2O2-dependent oxidases from
ligninolytic cultures of Phanerochaete chrysosporium. FEBS Lett. 169 (2), 247
250.
Lamed, R.J., Lobos, J.H., Su, T.M., 1988. Effects of stirring and hydrogen on
fermentation products of Clostridium thermocellum. Appl. Environ. Microbiol.
54 (5), 12161221.
Lawson, L.R., Still, C.N., 1957. The biological decomposition of lignin literature
survey. Tappi J. 40 (9), 56A80A.
Lee, J.-W., Jeffries, T.W., 2011. Efficiencies of acid catalysts in the hydrolysis of
lignocellulosic biomass over a range of combined severity factors. Bioresour.
Technol. 102 (10), 58845890.
Lynd, L.R., Weimer, P.J., van Zyl, W.H., Pretorius, I.S., 2002. Microbial cellulose
utilization: fundamentals and biotechnology. Microbiol. Mol. Biol. Rev. 66 (3),
506577.
Mason, W.H., 1926. Process and apparatus for disintegration of wood and the like.
US Patent 1,578,609.
McMillan, J.D., 1994. Pretreatment of lignocellulosic biomass. In: Himmel, M.E.,
Baker, J.O., Overend, R.P. (Eds.), Enzymatic Conversion of Biomass for Fuels
Production. American Chemical Society Washington, DC, pp. 292324.
Millett, M.A., Baker, A.J., Satter, L.D., 1976. Physical and chemical pretreatments for
enhancing cellulose saccharification. Biotechnol. Bioeng. Symp. 6, 125153
(United States) Medium: X.
Nakahara, Y., Yamauchi, K., Saka, S., 2014. MALDI-TOF/MS analyses of
decomposition behavior of beech xylan as treated by semi-flow hotcompressed water. J. Wood Sci. 60 (3), 225231.
Ogura, M., Phaiboonsilpa, N., Yamauchi, K., Saka, S., 2013. Two-step decomposition
behavior of rice straw as treated by semi-flow hot-compressed water (in
Japanese). J. Jpn. Inst. Energy 92 (5), 456.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

H. Rabemanolontsoa, S. Saka / Bioresource Technology xxx (2015) xxxxxx


Phaiboonsilpa, N., 2010. Chemical conversion of lignocellulosics as treated by twostep semi-flow hot-compressed water. In: Graduate School of Energy Science,
Doctoral dissertation, Kyoto University, Kyoto, Japan.
Plcido, J., Capareda, S., 2015. Ligninolytic enzymes: a biotechnological alternative
for bioethanol production. Bioresour. Bioprocess. 2 (1), 112.
Pollegioni, L., Tonin, F., Rosini, E., 2015. Lignin-degrading enzymes. FEBS J. 282 (7),
11901213.
Prawitwong, P., Waeonukul, R., Tachaapaikoon, C., Pason, P., Ratanakhanokchai, K.,
Deng, L., Sermsathanaswadi, J., Septiningrum, K., Mori, Y., Kosugi, A., 2013.
Direct glucose production from lignocellulose using Clostridium thermocellum
cultures supplemented with a thermostable beta-glucosidase. Biotechnol.
Biofuels 6 (1), 184.
Rabemanolontsoa, H., Kuninori, Y., Saka, S., 2015. High conversion efficiency of
Japanese cedar hydrolyzates into acetic acid by co-culture of Clostridium
thermoaceticum and Clostridium thermocellum. J. Chem. Technol. Biotechnol.
http://dx.doi.org/10.1002/jctb.4679.
Rabemanolontsoa, H., Saka, S., 2012. Characterization of Lake Biwa macrophytes in
their chemical composition. J. Jpn. Inst. Energy 91 (7), 621628.
Saha, B.C., Iten, L.B., Cotta, M.A., Wu, Y.V., 2005. Dilute acid pretreatment, enzymatic
saccharification and fermentation of wheat straw to ethanol. Process Biochem.
40 (12), 36933700.
Saka, S., 2001. Post-petrochemistry of woody biomass by supercritical water
treatment. Mokuzai Kougyou 56, 105110 (in Japanese).
Saka, S., 2009. Recent progress of biofuels in Japan. IEA Task 39 Newsletter, vol. 23.
International Energy Agency, pp. 210.
Salvachua, D., Karp, E.M., Nimlos, C.T., Vardon, D.R., Beckham, G.T., 2015. Towards
lignin consolidated bioprocessing: simultaneous lignin depolymerization and
product generation by bacteria. Green Chem. http://dx.doi.org/10.1039/
C5GC01165E.

Sasaki, M., Kabyemela, B., Malaluan, R., Hirose, S., Takeda, N., Adschiri, T., Arai, K.,
1998. Cellulose hydrolysis in subcritical and supercritical water. J. Supercrit.
Fluids 13 (13), 261268.
Srensen, H.R., Meyer, A.S., Pedersen, S., 2003. Enzymatic hydrolysis of watersoluble wheat arabinoxylan. 1. Synergy between a-L-arabinofuranosidases,
endo-1,4-b-xylanases, and b-xylosidase activities. Biotechnol. Bioeng. 81 (6),
726731.
Takada, M., Saka, S., 2015. Characterization of lignin-derived products from
Japanese cedar as treated by semi-flow hot-compressed water. J. Wood Sci.,
19
Tengerdy, R.P., Szakacs, G., 2003. Bioconversion of lignocellulose in solid substrate
fermentation. Biochem. Eng. J. 13 (23), 169179.
Trzcinski, A.P., Stuckey, D.C., 2015. Contribution of acetic acid to the hydrolysis of
lignocellulosic biomass under abiotic conditions. Bioresour. Technol. 185, 441
444.
Waksman, S.A., Cordon, T.C., 1939. Thermophilic decomposition of plant residues in
composts by pure and mixed cultures of microorganisms. Soil Sci. 47 (3), 217
226.
Watt, C., 1854. Improvement in the manufacture of paper from wood, US Patent
11343 A.
Wright, J.D., 1988. Ethanol from biomass by enzymatic hydrolysis. Chem. Eng. Prog.
84 (8), 6274 (United States) Medium: X.
Yoo, C.G., Nghiem, N.P., Hicks, K.B., Kim, T.H., 2011. Pretreatment of corn stover
using low-moisture anhydrous ammonia (LMAA) process. Bioresour. Technol.
102 (21), 1002810034.
Yoshida, H., 1883. LXIII.-Chemistry of lacquer (Urushi). Part I. communication from
the chemical society of Tokio. J. Chem. Soc. Trans. 43, 472486.
Yoshimura, M., Byrappa, K., 2008. Hydrothermal processing of materials: past,
present and future. J. Mater. Sci. 43 (7), 20852103.

Please cite this article in press as: Rabemanolontsoa, H., Saka, S. Various pretreatments of lignocellulosics. Bioresour. Technol. (2015), http://dx.doi.org/
10.1016/j.biortech.2015.08.029

Вам также может понравиться