You are on page 1of 10

Industrial Crops and Products 71 (2015) 163–172

Contents lists available at ScienceDirect

Industrial Crops and Products


journal homepage: www.elsevier.com/locate/indcrop

Oxalic acid as a catalyst for the hydrolysis of sisal pulp


Talita M. Lacerda, Márcia D. Zambon, Elisabete Frollini ∗
Macromolecular Materials and Lignocellulosic Fibers Group, Center for Science and Technology of BioResources, Institute of Chemistry of São Carlos,
University of São Paulo, CP 780, 13560-970 São Carlos, São Paulo, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: The possible shortage of crude oil has led to the search for alternative sources of chemicals and energy
Received 8 May 2014 producers from biomass.The current results are related to the analysis of the unreacted cellulosic material
Received in revised form 18 February 2015 and the liquor, from the acid hydrolysis of an industrial Kraft sisal pulp. An important characteristic
Accepted 22 March 2015
involving the acid hydrolysis of biomass is the possibility to utilize different acids because the only
requirement for the reaction to occur is the presence of a proton source in the media. In this context,
Keywords:
a series of acid hydrolysis reactions of a prior mercerized sisal pulp was performed with 4.6 mol L−1
Sisal pulp
oxalic acid at varying temperatures (from 80 to 100 ◦ C), and aliquots were withdrawn from the reaction
Acid hydrolysis
Oxalic acid
media.The liquor composition was analyzed by HPLC and the unreacted pulps (e.g., the material that
was not decomposed to sugars) were characterized by SEM, average molar mass, crystallinity index, and
their size distribution was determined using a fiber analyzer. The kinetic model developed by Saeman
was applied to calculate the activation energy for the acid hydrolysis of sisal pulp. A higher glucose
concentration (approximately, 8 g L−1 ) was detected at 100 ◦ C, after 16 h. The percentage of hydrolyzed
cellulose under these conditions was 50.4%, and the amount of glucose recovered as a percentage of
the amount in the raw sisal pulp was 22.3%. Unreacted cellulose was isolated at a low average molar
mass (approximately 6000 g mol−1 ) and high crystallinity (approximately 88%), at 100 ◦ C. Oxalic acid is
a promising catalyst for the hydrolysis of cellulose for the production of sugars and for pretreatment of
the fibers prior to other processes (e.g., enzymatic hydrolyses).
© 2015 Elsevier B.V. All rights reserved.

1. Introduction the following processes are required in this order: concentration,


detoxification, neutralization and supplementation with nutrients
The current trend toward environmentally friendly energy has (Lenihan et al., 2010).
promoted a worldwide effort in recent decades to develop the use Biomass obtained from different raw materials, including corn-
of biomass for the productionof biofuels so that it is economi- cob (Sun et al., 2013), rice hulls (Dagnino et al., 2013), palm leaves
cally feasible (Lu and Mosier, 2008; Tomme et al., 1995; Xiang (Werle et al., 2013), sugarcane bagasse (Rocha et al., 2012), and
et al., 2004; Xu et al., 2011). Therefore, ethanol has become the softwood pulps (Hoeger et al., 2013), have been investigated as a
most promising short-term alternative fuel due to its compatibil- source of sugar for further conversion to ethanol. When a lignocel-
ity with current car engines and the gasoline supply chain (Cole, lulosic feedstock is used as a source of sugar in ethanol production,
2011). In this context, the search for non-fossil fuels has led to the the polysaccharides content, especially the cellulose content, of the
development of so-called second-generation ethanol (e.g., ethanol respective feedstock is highly relevant. In the present study, sisal
derived from biomass) through the acid or enzymatic saccharifica- was chosen as the lignocellulosic raw material due to the high cel-
tion. lulose content of its fibers (approximately 65% cellulose (Da Silva
The possibility of converting cellulose into glucose is especially et al., 2012; Frollini et al., 2013) and other reasons, as its abundance
useful because these sugars can be easily converted to ethanol, in Brazil (de Paula et al., 2012).
which justifies the growing interest in research into the use of The production of bioethanol from lignocellulosic biomass
lignocellulosic material, which is typically agricultural waste, for typically requires some sort of pretreatment to disrupt the
the production of bioethanol. The hydrolysates obtained from the cellulose–hemicelluloses–lignin complex to make the cellulose and
hydrolysis are processed for use as fermentation media. In general, hemicellulose more accessible for conversion into sugars. Diverse
pretreatments have been reported in the literature, such as steam
explosion (Rocha et al., 2012; Liu et al., 2013; Martin-Sampedro
∗ Corresponding author. Tel.: +55 16 3373 9951; fax: +55 16 3373 9952. et al., 2014; Rocha et al., 2014), microwave treatment (Xu et al.,
E-mail address: elisabete@iqsc.usp.br (E. Frollini). 2011), mechanical fibrillation (Hoeger et al., 2013), organosolv (Koo

http://dx.doi.org/10.1016/j.indcrop.2015.03.072
0926-6690/© 2015 Elsevier B.V. All rights reserved.
164 T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172

et al., 2012), and treatments with concentrated phosphoric acid and catalyst. This process exhibited good performance with a low yield
ionic liquids (Sathitsuksanoh et al., 2012). The cellulosic pulp or cel- of decomposition products.
lulose obtained from the pretreatment of lignocellulosic biomass is Although a wide range of acids may be used as a source of pro-
often further subjected to treatments that disrupt the structure of tons for cellulose hydrolysis, the most commonly used is sulfuric
cellulose in order to increase the accessibility of the hydrolyzing acid, which is the only common polyprotic acid that exhibits com-
agents (Abushammala and Hashaikeh, 2011). The pretreatments plete first deprotonation. The second deprotonation only slightly
may have two functions, i.e., to release the cellulose from the resid- increases the molarity of H3 O+ (i.e., pKa1 = −3 and pKa2 = 1.92 at
ual recalcitrant lignin and disrupt its crystalline structure, as is the 25 ◦ C) (Budavari, 1996). Oxalic acid, which is a dicarboxylic acid
case when alkali pretreatment is utilized (Rawat et al., 2013). In the (HOOC-COOH), shows a lower degree of ionization than sulfuric
present study, the sisal lignocellulosic fiber was delignified on an acid (pKa1 = 1.27 and pKa2 = 4.28 at 25 ◦ C (Budavari, 1996) and may
industrial scale through a Kraft process, as described in Section 2.1, require more severe conditions for its use in cellulose hydrolysis.
and the cellulosic pulp received from the industrial process was However, this acid has the advantage that it can be obtained from
treated with an aqueous NaOH solution via mercerization (Lacerda natural sources, such as biomass. Fermentation and nitric acid oxi-
et al., 2012; Okano and Sarko, 1985). dation of carbohydrates can be used for the production of oxalic acid
The mercerization process was used for pretreatment of the sisal (Gürü et al., 2001; Sullivan et al., 1983). Oxalic acid, which is present
pulp, which was determined to be efficient in a previous study in many plants and vegetables especially in the Rumex and Oxalis
(Lacerda et al., 2012). This treatment alters the fine structure and families, is also a metabolic product in many fungi. Some species of
morphology of the fiber as well as the conformation of the cellulose Penicillium and Aspergillus convert sugar into calcium oxalate with
chains (transforming it from cellulose I to II). During this process, yields up to 90% (Budavari, 1996).
the material swells, the polysaccharide chains are rearranged, and In recent years, oxalic acid has been used for the pretreatment
usually the crystalline portion decreases. These changes result in of lignocellulosic biomass (Hong et al., 2013), especially prior to
higher adsorption because mercerization increases the specific sur- enzymatic hydrolysis. In this application, oxalic acid shows advan-
face area of the fiber, thus making the chains more accessible to tages over sulfuric acid because it is less toxic to yeasts and other
reagents (Ass et al., 2006; Dinand et al., 2002; Okano and Sarko, microbes and does not produce prejudicial smells (Bo et al., 2013;
1985). In general, the crystallinity index of cellulose fibers has a Lee et al., 2011). In addition, oxalic acid can be recovered more
direct relationship to reagent accessibility (Ciolacu et al., 2012), easily than sulfuric acid via techniques conventionally used for
and an increase in the crystallinity index decreases the hydroly- organic acids, such as ion-exchange and adsorption, or via less con-
sis ratio and sugar production (Wang et al., 2012). Therefore, prior ventional techniques, such as electrodialysis, which can offset the
to their hydrolysis, the lignocellulosic sisal fibers were subjected higher cost of oxalic acid on a weight basis compared to sulfuric
first to Kraft process and then to mercerization. The combination acid (Lee and Jeffries, 2013). Additionally, vom Stein et al. (2011)
of two pretreatments (as two different steps or as a coupled step) have shown that after using oxalic acid to hydrolyze hemicelluloses
prior to the hydrolysis of the cellulosic biomass has been consid- and release lignin from a lignocellulosic source, when the aque-
ered in several studies, as for instance pretreatment with alkaline ous phase was concentrated oxalic acid precipitated as a pure solid
peroxides followed by steam explosion (Balcu et al., 2011; Chen (approximately 85 wt%) and could be recovered by simple filtra-
and Qiu, 2010), ionic liquid treatment coupled with steam explo- tion. The purity of the recovered oxalic acid allowed its direct reuse
sion (Chen and Qiu, 2010), alkali combined extrusion pretreatment as a catalyst. This set of positive aspects, among others, made the
(Zhang et al., 2012), among others. use of this acid in sustainable biorefineries particularly attractive
Over the years, a number of methods for the conversion of (Hongsiri et al., 2015)
cellulose to glucose involving the depolymerization of cellulose In this study, which is a continuation of a previous study that
via hydrolysis have been developed (Arantes and Saddler, 2010; utilized sulfuric acid (de Paula et al., 2012; Lacerda et al., 2012;
Bustos et al., 2003; Cho et al., 2011; Martín and Grossmann 2012; Lacerda et al., 2013), industrial cellulosic sisal pulp was used as the
Rodrigues and Guirardello, 2008; Shields and Boopathy, 2011; starting material. Thus, the focus of the present work was not to use
Tásic et al., 2009). Among these methods, the two most commonly oxalic acid as a pretreatment to remove lignin and hemicellulose,
used involve cellulolytic enzymes (Arantes and Saddler, 2010; namely to produce a pulp prior to saccharification (Lee et al., 2011).
Dagnino et al., 2013; Hoeger et al., 2013) or acids, including various The main goal was to explore the performance of this acid as a
strengths of sulfuric acid (de Paula et al., 2012; Lacerda et al., catalyst for the saccharification of cellulosic pulp.
2012). Solid acid catalysts have also been employed (Ma et al., The hydrolysis reactions were conducted with mercerized sisal
2012; Zhang et al., 2013). pulp and monitored over time by evaluating the liquor (sugars and
Enzymes are still relatively expensive for application to the their decomposition products), which is an important stage in the
large-scale production of fuel ethanol from biomass. Therefore, acid production of bioethanol. The physicochemical and morphological
catalysis is more attractive even though acids tend to degrade glu- characteristics of the solid non-hydrolyzed material were deter-
cose. This degradation lowers the yield of fermentable sugars from mined to provide important information on the overall process. In
the biomass and results in degradation products, such as hydrox- addition, these results are relevant for applications in the materials
ymethyl furfural, levulinic acid, and formic acid, which inhibit yeast area, such as for the preparation of crystalline cellulosic materials.
fermentation (Mosier et al., 2001). To complement the investigation on the performance of oxalic
An interesting feature that involves the acid hydrolysis of acid as a catalyst, the kinetic model developed by Saeman et al.
biomass is the possibility of using different acids as catalysts (1945), which is based on the acid hydrolysis reactions of wood
because in principle, the process only requires a source of protons at low acid concentrations and high temperatures, was applied to
in the aqueous medium for the reaction to occur. Several studies calculate the activation energy for the acid hydrolysis of sisal pulp.
have been conducted involving various types of acids (Sivers and
Zacchi, 1995; Taherzadeh and Karimi, 2007), including H2 SO4 (de 2. Materials and methods
Paula et al., 2012; Lacerda et al., 2012; Lacerda et al., 2013), HCl
(Hernández-Salas et al., 2009), H3 PO4 (Gámez et al., 2004; Lenihan 2.1. Materials
et al., 2010), heteropoly acids (Tian et al., 2010), and formic acid
(Sun et al., 2008). Marzialetti et al. (2011) described a process The sisal pulp used was obtained from the leaves of plants
involving dilute hydrolysis of switch grass using formic acid as the cultivated in the state of Bahia, a semi-arid region of North-
T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172 165

east Brazil (daily average temperature: exceeds 24 ◦ C, highest crystallinity index (CI), scanning electron microscopy (SEM), and
altitude: 600 m, average annual rainfall: from 650 to 1250 mm, the average thickness and length of the fibers, as described below
average annual relative air humidity: from 66 to 79%, winds: 3 ms−1
(http://sistemasdeproducao.cnptia.embrapa.br/FontesHTML/Sisal/). 2.2.3. Pulp characterization
The sisal pulp used in all stages of this study was kindly provided Initial and non-hydrolyzed pulps, which were withdrawn dur-
by Lwarcell (Lençóis Paulista, São Paulo, Brazil), where the ligno- ing the reactions, were characterized as described in the following
cellulosic fiber was delignified via the Kraft process and bleached sections.
with Elementary Chlorine Free (ECF) sequences to produce fibers
that averaged 2.9 mm in length and 18.4 ␮m in thickness with an 2.2.3.1. Hemicellulose and ˛-cellulose content. The procedure con-
86% ISO brightness, as reported by the supplier. Klason lignin was sists of adding an amount of pulp to a 17.5% NaOH solution (liquor
analyzed according to TAPPI T222-om02, and no insoluble lignin ratio = 1:20, w/v), at room temperature (Sun et al., 1995). The ␣-
was detected. All of the pulp samples were previously mercerized cellulose content was determined from the dry masses (dried in an
(as described in Section 2.2.1), milled in a MARCONI MA048 miller air-circulating stove at 105 ◦ C, up to constant weight) of pulp before
(30 mesh stainless steel) and dried in a vacuum oven at 60 ◦ C for and after the treatment with the NaOH solution. The difference
4 h. between the initial mass and the calculated mass of ␣-cellulose
The pulp was mercerized in a sodium hydroxide solution corresponds mainly to the hemicellulose content (Browning, 1967).
(NaOH in pellets, Qhemis) (ASTM1695-07). For the hydrolysis reac-
tions, oxalic acid (Cromato) was used. A 2% cupri-ethylenediamine 2.2.3.2. Scanning electron microscopy (SEM). For the SEM measure-
(CED) solution (Qeel) was used in the viscometric analysis. Glu- ments, a Zeiss LEO-440 instrument was used with a tungsten
cose, xylose, arabinose, formic acid, acetic acid, furfural and filament to generate electrons, as previously described (Ramires
hydroxymethylfurfural (Sigma–Aldrich) were used as standards for et al., 2010). The samples were dried in a vacuum for 4 h and covered
high-performance liquid chromatography (HPLC). with gold.

2.2.3.3. Average molar mass (MMvis) by viscometry. Viscometry was


2.2. Methods used to determine the average molar mass of the starting and non-
hydrolyzed cellulose using aliquots withdrawn from the reaction
2.2.1. Mercerization medium at different time intervals during the reactions. Cellulose
The process was performed with 20 g of pulp suspended in 1 L was dissolved in a copper(II)-ethylenediamine/water solution (1:1,
of a 20 wt% NaOH solution for 3 h at 50 ◦ C with stirring. The pulp v/v), and the flow time in a glass capillary viscometer (Ubbelohde,
was filtered, washed until the washing water reached its initial pH f = 0.63 mm, AVS-350 Schott-Geräte) was measured. The average
(approximately 6.5), and dried at room temperature and then in a molar mass of the samples was determined according to TAPPI T230
vacuum oven at 60 ◦ C for 4 h. These conditions were chosen based om-2008, as previously described (Morgado et al., 2011). All of the
on the results obtained in a prior study (Lacerda et al., 2012). measurements were performed in triplicate.

2.2.3.4. Crystallinity index (CI) by X Ray diffraction. Initial and non-


2.2.2. Acid hydrolysis hydrolyzed cellulose (obtained from five pulp samples, t = 1, 4, 8,
The reactions were performed under mechanical stirring, in a 12 and 16 h of reaction), previously dried in a vacuum oven (60 ◦ C,
jacketed reactor connected to a thermostatic bath, and provided 4 h), were analyzed by X-ray diffraction to determine the CI using
with a condenser and a connection to withdraw aliquots. For the a universal X-ray diffractometer (URD-6 model, CARL ZEISS JENA,
hydrolysis reactions, 900 mL of a 4.6 mol L−1 oxalic acid solution 40 kV/20 mA,  (Cuk␣) = 1.5406 Å). The CI was calculated according
were heated (80, 85, 90, 95 and 100 ◦ C), and 30 g of dried, milled to the following equation:
and mercerized pulp was added. All of the reactions were per-
I1
formed over a period of 16 h.The conditions were chosen such that CI = 1 − (1)
the results obtained in this study could be compared to those of I2
previous studies (de Paula et al., 2012; Lacerda et al., 2012; Lacerda where I1 is the minimum intensity, which is proportional to the
et al., 2013). The initial results for the glucose yield indicated that non-crystalline fraction of the cellulose (18◦ ≤ 2 ≤ 19◦ for cellu-
the length of time should be longer than that used in previous stud- lose I and 13◦ ≤ 2 ≤ 15◦ for cellulose II), and I2 is the maximum
ies (i.e., 6 h) to obtain similar yields. Therefore, a time interval of intensity, which corresponds to the signal of the crystalline frac-
16 h was used in the current study. tion (22◦ ≤ 2 ≤ 23◦ for cellulose I and 18◦ ≤ 2 ≤ 22 ◦ for cellulose
During the reaction course, aliquots of the suspension were II) (Buschle-Diller and Zeronian, 1992).
withdrawn (t = 1, 2, 3, 4, 5, 6, 8, 10, 12, 14 and 16 h), and the solid (the Because the diffractogram is the intensity as a function of the
unreacted, non-hydrolyzed pulp) was separated from the liquor by diffraction angle (2), it is possible to obtain the half-height [(2)]
simple filtration. The liquor was stored at approximately 5 ◦ C for for the maximum diffraction intensity. Using the Scherrer equation
further analysis. The sugars and related decomposition products (Pouget et al., 1991), the average size of the crystalline domains can
were analyzed by high-performance liquid chromatography (HPLC) be calculated:
using a Shimadzu instrument. Glucose, xylose, arabinose, formic 0.9 × 
acid and acetic acid were analyzed using a refractive index detec- L= (2)
(2) × cos
tor (RID-6A; Aminex HPX 87H column (300 × 7.8 mm BIO-RAD);
eluent: 0.005H2 SO4 mol L−1 ; flow: 0.6 mL min−1 ; 45 ◦ C). Furfural where  = 1.5406 Å, which is the wavelength of the X-rays
and hydroxymethyl furfural were analyzed using a UV detector at employed.
274 nm (SPD-10AV); BondapakTM C18 column (Waters); eluent:
acetonitrile/water 1:8 (v/v) acetic acid 1% (v/v); flow: 0.8 mL min−1 ; 2.2.3.5. Average thickness and length of the fibers. The average thick-
25 ◦ C). All HPLC analyses were performed in duplicate. ness and length of the fibers were obtained using MorFi Compact
The non-hydrolyzed pulp was thoroughly washed, stored in (Techpap) equipment, which measures approximately 5000 fibers
Eppendorf tubes and dried in a vacuum oven at 60 ◦ C for 4 h prior suspended in water. The MorFi automatic fiber-analysis system also
to determination of the viscometric average molar mass (MMvis), provides the fiber-occurrence density (Millions-per-Gram), which
166 T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172

Fig. 1. SEM images for the starting pulp (a) and non-hydrolyzed pulps from the reactions performed at 80 ◦ C (b), (e), 90 ◦ C (c, f) and 100 ◦ C (d, g) after 6 h (b–d) and 16 h (e–g).

enables the construction of a density map, i.e., a three-dimensional comparison between the results reported by Lacerda et al., 2012
graph with the following characteristics: average length values on (hydrolysis of mercerized sisal pulp) and those found in a previous
the x axis, average thickness values on the y axis, and the fiber- study on the hydrolysis reactions catalyzed by sulfuric acid solu-
occurrence density values on the z axis. tions using unmercerized sisal pulp (de Paula et al., 2012) showed
that the changes caused by mercerization significantly improved
3. Results and discussion the accessibility of the cellulose chains to the reagents. Under the
same reaction conditions (30% sulfuric acid, 100 ◦ C, 6 h), the hydrol-
3.1. Initial pulp characterization ysis of mercerized pulp generated yields of up to 50% more glucose,
compared to the unmercerized pulp.
The values obtained for the average molar mass (MMvis),
␣-cellulose content, hemicellulose content and CI for the pulp
samples were previously presented (Lacerda et al., 2012) and are
119000 ± 700 g mol−1 , 88 ± 4%, 12 ± 3% and 77%, respectively, for 3.2. Characterization of the non-hydrolyzed pulp
the unmercerized pulp and 99,900 ± 450 g mol−1 , 98 ± 1%, 2 ± 1%
and 69%, respectively, for the mercerized pulp. 3.2.1. SEM
As previously discussed (Lacerda et al., 2012), the merceriza- Fig. 1 shows the SEM images of the initial pulp used as raw
tion pretreatment is important for the acid hydrolysis of sisal pulp material for the hydrolysis reactions (a) and the non-hydrolyzed
because it produces a material with a higher ␣-cellulose con- cellulose, which was withdrawn at different time intervals (6 and
tent compared to the untreated pulp, thus resulting in fewer side 16 h) and temperatures (80, 90 and 100 ◦ C) (b–g) during the hydrol-
reactions and undesirable byproducts in the reaction medium. In ysis reaction.
addition, the alkaline treatment led to the formation of cellulose At 100 ◦ C, greater fragmentation of the fibers occurred com-
II, as indicated by the characteristic peaks in the diffractograms pared to the non-hydrolyzed fibers at 80 ◦ C (Fig. 1d–g and b–e,
with 2 between 18◦ and 22◦ (Lacerda et al., 2012) and reduced the respectively). In addition, longer reaction times complement this
CI value. Ciolacu et al. (2012) reported that the allomorph with the effect, which can be observed by comparing images, c and f (6 and
most accessible crystal surfaces and amorphous regions is cellulose 16 h of reaction, respectively, 90 ◦ C) and images d and g (6 and 16 h
II. The increased amount of chains in the non-crystalline regions, of reaction, respectively,100 ◦ C). These images represent the aver-
which are more accessible to reagents, and the change of cellu- age of all ofthe other reaction conditions, which resulted in the
lose I to II due to mercerization favor the hydrolysis process. A same pattern of fiber fragmentation observed by SEM.
T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172 167

For the shorter reaction times, the average size of the crystalline
domains increased at all of the temperatures considered in this
study (Fig. 2c). This result may be due to the joining of smaller
crystalline domains, due to the consumption, during hydrolysis, of
chains present in non-crystalline domains and that were initially
allocated among the smaller crystalline domains. When the tem-
perature is increased to 100 ◦ C, an initial increase in the dimensions
of the crystallites also occurs (Fig. 2c), followed by a reduction after
4 h. This phenomenon is accentuated after 12 h, which is most likely
due to the hydrolysis of chains in the crystalline domain, which also
leads to a decrease in CI (Fig. 2b).

3.2.3. Average thickness and length of fibers (Fiber analyzer –


MorFi).
Fig. 3 shows the density maps obtained from the morphological
analysis using the MorFi Compact equipment for the initial pulp
(mercerized), (Lacerda et al., 2013) and the non-hydrolyzed cellu-
lose withdrawn from the reaction media at 80, 90 and 100 ◦ C after
6 and 16 h of hydrolysis. The reactions at other temperatures (i.e.,
85 and 95 ◦ C, data not shown) exhibited similar behavior.
All of the fibers withdrawn from the reaction media (non-
hydrolyzed pulp) exhibited severe variations in their length and
thickness. From Fig. 3b, which corresponds to 6 h at 80 ◦ C,the
density map exhibits higher quantities of fibers with a length of
129–215 ␮m compared to the initial pulp (Fig. 3a), (Lacerda et al.,
2013) thus indicating that a greater population of fibers have a nar-
rower size distribution of lengths after hydrolysis. The variation in
the fiber thickness after hydrolysis is not as clear as the variation in
the fiber length. However, thinner fibers were formed as the tem-
perature and the exposure time to the acid solution were increased,
which is evident from the increase in the density of the fibers with a
thickness of 11–14 ␮m (Fig. 3b) at higher temperatures and longer
reaction times.
By comparing Fig. 3b and c with Fig. 3f and g, an increase in
the density of the less thick [11–23 ␮m] and shorter [46–129 ␮m]
fibers was observed, and this increase is more significant when
T = 100 ◦ C (Fig. 3f and g) compared to T = 80 ◦ C (Fig. 3b and c). For
the initial pulp (Fig. 3a), (Lacerda et al., 2013) the fraction with a
Fig. 2. (a) Average molar mass, (b) crystallinity index and (c) crystalline domain greater density is located in a broader thickness region [11–39 ␮m]
dimensions for the non-hydrolyzed cellulose during the hydrolysis reactions. compared to the maps of the non-hydrolyzed pulps (Fig. 3b–g).
All of these maps exhibit a “valley” dividing two regions of differ-
ent thicknesses, which is due to the consumption of the fractions
3.2.2. The average molar mass (MMvis), crystallinity index (CI) between [6 and 8] and [11 and 14] ␮m during the hydrolysis
and crystalline domain dimensions via X-ray diffraction. reactions.When sulfuric acid was used at the same temperature
Fig. 2 shows the average molar mass (a), crystallinity index (b) and concentration (Lacerda et al., 2013), the corresponding den-
and average size of the 0 0 2 plane (c) as a function of time for the sity map for the 6 h reaction at 100 ◦ C exhibited an even higher
hydrolysis reactions. fiber density in regions with more narrow [11–14 ␮m] and shorter
In Fig. 2a, the drastic reduction in MMvis, which was observed [46–77 ␮m] fibers, which indicates that oxalic acid can be uti-
at all of the temperatures, indicates that the oxalic acid acted lized for cellulose hydrolysis under milder conditions than with
as a catalyst for the hydrolysis of sisal pulp. The initial MMvis sulfuric acid.
value (99,900 g mol−1 ) was reduced after 16 h of reaction to val- The density maps shown here represent the average of all of the
ues ranging from approximately 6000–7000 g mol−1 depending on other reaction conditions at intermediate time and temperature,
the temperature (Fig. 2a). which showed similar patterns for the evolution of the length and
The preferential hydrolysis of the chains in the non-crystalline thicknessof the fibers.
domains led to a significant increase in CI since the early reaction
times (Fig. 2b). The substantial decrease in the average molecular 3.3. Characterization of the products of the hydrolysis reactions
weight led to significantly shorter chains compared to the starting
pulp used in the hydrolysis, which can favor the organization of the As described in Section 2.2.2, the sugars and their decomposition
chains, which can contribute to the increase in CI. A decrease in the products analyzed in the liquor were glucose, xylose, arabinose,
average molar mass, as well as an increase in CI compared to the formic acid, acetic acid, hydroxymethylfurfural (HMF) and furfural.
beginning of the reaction, were also observed in studies where sul- Glucose (produced primarily by the hydrolysis of cellulose), xylose
furic acid was used as the catalyst for the hydrolysis of sisal pulp (de (produced by the hydrolysis of hemicellulose), HMF and furfural,
Paula et al., 2012; Lacerda et al., 2012). It is important to note that which are the main products from the decomposition of glucose
higher temperatures and longer reaction times increase the hydrol- and xylose, respectively, were observed in all of the reactions. The
ysis of the chains in the crystalline regions, leading to a decrease in presence of glucose under all hydrolysis conditions indicates cleav-
CI (Fig. 2b). age of the 1,4-␤-glycosidic linkages of the cellulosic chains, and the
168 T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172

Fig. 3. Fiber-occurrence density (Millions-per-Gram) of the initial pulp (a, (Lacerda et al., 2013)) and non-hydrolyzed cellulose withdrawn from the reaction performed at
80 ◦ C (b and c), 90 ◦ C (d and e) and 100 ◦ C (f and g) after 6 h and 16 h, respectively.
T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172 169

Glucose production is directly related to the temperature of the


reaction, and the maximum concentration detected was approxi-
mately 8 g L−1 (after 16 h, 100 ◦ C, Fig. 5a). The amount of glucose in
the raw sisal pulp was 35.9 g L−1 {[glucose] sisal pulp = mass (30 g) × %
cellulose (98%) × q (1.1)/V (0.9L)}, where q is the conversion factor
for anhydrous glucose (162 g mol−1 , in cellulose chain) to glucose
in the liquor (180 g mol−1 , they differ in mass by a water molecule).
Thus, the amount of glucose recovered after 16 h at 100 ◦ C was
22.3% (8/35.9 × 100).
At 80 ◦ C and 85 ◦ C, almost no HMF was formed, and only after
10 h a small fraction was detected at 85 ◦ C. For the reaction per-
formed at 90 ◦ C at reaction times up to 6 h, the production of HMF
is absent or very small, and after 8 h, less HMF was produced com-
pared to that produced at 95 ◦ C. At 80 ◦ C and at reaction time of
16 h, approximately 1.3/0.005 g L−1 of glucose/HMF was observed
in the medium, and at 100 ◦ C, approximately 8.0/0.092 g L−1 of glu-
cose/HMF was observed (Fig. 5a and b).
At 80 ◦ C and 85 ◦ C, the concentration of formic acid was zero
or quite low, and at 90 ◦ C, the production of formic acid followed
the trend observed for HMF. For the reactions performed at 95 to
100 ◦ C, the production of formic acid was observed practically from
the beginning of the reactions(Fig. 5b and c) due to the extremely
low stability of HMF under these conditions, which leads to the
formation of formic and levulinic acids (not analyzed).
Fig. 4. (a) Xylose (Errors from ±0.01 to ±0.22, and (b) furfural (Errors from ±0.007 The amount of glucose present in the liquor (8 g L−1 , Fig. 5a)
to ±0.25) concentrations as function of hydrolysis time. together with the amount of its decomposition products at 100 ◦ C
after 16 h (Fig. 5b and c), was 18.1 g L−1 , which means that the
percentage of hydrolyzed cellulose corresponded to 50.4% (18.1 /
presence of HMF indicates that all of the reaction conditions led to 35.9 × 100).
the partial decomposition of glucose by dehydration. Formic acid,
which is a decomposition product of HMF (Belgacem and Gandini, 3.3.1. Activation energy for acid hydrolysis.
2008), was detected at all of the temperatures studied. Arabinose As previously described (de Paula et al., 2012), to calculate the
and acetic acid, which are both produced from the hydrolysis of reaction constants (i.e., k1 and k2, related to the rates of pulp hydrol-
hemicelluloses, were not observed for any of the reactions con- ysis and decomposition, respectively), the Saeman model (Saeman
sidered here. During kraft cooking the acetyl groups are readily et al., 1945) was applied to the experimental glucose concentration
hydrolysed (Sjöström, 1993), which may explain the absence of curves using the Origin Pro 8 SRO software version 8.0724 (B724).
acetic acid, since a deacetylated xylan was present in the starting With the nonlinear fitting tool software, the data were fitted to the
pulp. model. For this purpose, the values of the glucose concentration
Fig. 4 shows the products observed in the liquor that are directly ([B]) were inserted into the equation that describes the model (Eq.
obtained from hemicellulose hydrolysis (i.e., xylose) (Fig. 4a) and (4)) for each time (t) starting from the initial cellulose concentra-
its major decomposition product, furfural (Fig. 4b). tion ([A]0 ), and the values of k1 and k2 were calculated for each
For the reactions performed at 80, 85, 90 and 95 ◦ C, the concen- of the studied conditions. The concentration of the initial cellulose
tration of xylose increased until a reaction time of approximately was calculated based on the mass of the pulp added to the reactor
12 h (Fig. 4a), and the concentration then remained nearly constant and the ␣-cellulose content (98%) of the starting mercerized pulp.
at longer reaction times. In addition, at 100 ◦ C, the concentration of
k1
xylose increased until a reaction time of approximately 6 h followed [B] = (e−k1 t − e−k2 t )[A]0 (4)
k2 − k1
by a decrease in the observed concentration, which indicates that
at this temperature, the decomposition of xylose was significantly Fig. 6 shows the nonlinear fit of the Saeman model for the glu-
more intense than at the other temperature conditions. All of the cose concentration as a function of the hydrolysis time (90 ◦ C). The
reaction conditions led to the decomposition of xylose, which was values of the constants (k1 and k2 ) and the R2 value are shown in
confirmed by the presence of furfural at all of the studied tempera- Table 1. The profile of the other reactions (80,85, 95, and 100 ◦ C,
tures, as shown in Fig. 4b. However, from 80 ◦ C to 90 ◦ C for reaction figures not shown) was close to the model proposed by Saeman
times up to 8 h, the concentration of furfural was zero or quite (Saeman et al., 1945), as indicated by the values of R2 (Table 1).
low. In addition, at 80 ◦ C and 16 h, approximately 0.6/0.05 g L−1 of Oxalic acid does not catalyze the decomposition of glucose at
xylose/furfural was observed in the medium, and at 100 ◦ C, approx- temperatures less than 90 ◦ C. At 95 ◦ C and 100 ◦ C, k2 is not zero, but
imately 1.5/0.5 g L−1 of xylose/furfural was observed (Fig. 4a and k1 assumes values that are considerably higher than those at lower
b). temperatures (Table 1).
Fig. 5 shows the products found in the liquor that are primarily In a prior study, the reactions were performed with 4.6 mol L−1
obtained from cellulose hydrolysis (i.e., glucose, Fig. 5a), its major sulfuric acid (Lacerda et al., 2012), which is the same concentra-
decomposition product (i.e., HMF, Fig 5b), and formic acid (Fig. 5c). tion used here for oxalic acid, and at 100 ◦ C,which is the highest
In sisal pulp, hemicellulose contains only approximately 2.2% glu- temperature tested for oxalic acid. It should be noted that when
cose (Megiatto et al., 2007). As mentioned above, the content of sulfuric acid was used as the catalyst, aliquots were withdrawn
␣-cellulose in the mercerized sisal pulp used in the current study is over a period of 6 h, and for oxalic acid, the first aliquot was
98% (i.e., the content of hemicellulose is approximately 2%). There- obtained at 1 h and more aliquots were taken until 16 h (see Sec-
fore, as an approximation, the glucose content was considered to tion, 2.2). In addition, although the concentration of both acids
originate only from cellulose. (i.e., sulfuric and oxalic) was the same, the concentration of pro-
170 T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172

Fig. 5. (a) Glucose (Errors from ±0.01 to ±0.69), (b) HMF (Errors from ±0.01 to ±0.29) and (c) formic acid (Errors from ±0.13 to ±0.66) concentrations as function of the
hydrolysis time.

Table 1
Constants for pulp hydrolysis (k1 ) and glucose decomposition (k2 ) and R2 values for the oxalic acid catalyst.

80 ◦ C (min−1 ) 85 ◦ C (min−1 ) 90 ◦ C (min−1 ) 95 ◦ C (min−1 ) 100 ◦ C (min−1 )

k1 x 103 1.0 4.7 7.4 21.4 33.1


k2 x 103 0.0 0.0 0.0 4.6 6.4
R2 0.996 0.983 0.992 0.986 0.972

tons was considerably higher for sulfuric acid due to the differences reaction is slower than the reaction under sulfuric acid catalysis at
in their respective ionization potentials. The results for k1 and k2 the same temperature. The relatively weaker ionization potential of
obtained were 5.5 10−3 min−1 and 6.6 10−3 min−1 (reaction time: oxalic acid compared to sulfuric acid could decrease the occurrence
6 h), respectively. For oxalic acid, in the current study the values of of subsequent dehydration reactions (Lee et al., 2011).
these constants were 33.1 10−3 min−1 and 6.4 10−3 min−1 (reaction The values obtained using sisal pulp and oxalic acid at 100 ◦ C
time: 16 h) (Table 1). Therefore, despite the differences between the have the similar magnitudes, more so for k1 , as those presented
two reactions, it can be inferred that under oxalic acid catalysis, at a by Aguilar et al. (2002), who obtained values for k1 and k2 of 35.7
temperature at which glucose decomposition occurs (100 ◦ C), this 10−3 min−1 and 0.29 10−3 min−1 , respectively, for the hydrolysis
T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172 171

to crystallinity, among others, were not present, contrary to what


occurs when pulps are hydrolyzed, as in this study.

4. Conclusions

Under the conditions studied, the results indicated that oxalic


acid can be used as a catalyst for the acid hydrolysis of sisal pulp as
a first step in the production of second generation ethanol because
it allows the formation of glucose even at the lowest temperature
evaluated (80 ◦ C). Higher temperatures facilitate the process, and a
higher glucose concentration (approximately 8 g L−1 ) was detected
when T = 100 ◦ C, which can be considered as the optimal reaction
condition of the present study concerning glucose production.
Unreacted cellulose (e.g., a material that was not totally decom-
posed to simple sugars) was isolated at a low average molar
mass (approximately 6000 g mol−1 , when T = 100 ◦ C) and high crys-
tallinity index (approximately 88%, when T = 100 ◦ C), which are
characteristics of a low DP microcrystalline cellulose. Under mild
Fig. 6. Nonlinear adjustment of the Saeman kinetic model for the reaction at 90 ◦ C
conditions (i.e., an acid concentration and a reaction time of
with 4.6 mol L−1 oxalic acid.
0.9 mol L−1 and 6 h, respectively), oxalic acid can also be used as
a pretreatment agent for sisal pulp because it eliminates approx-
imately 42% of the hemicellulose present in the pulp, which is
responsible for the generation of undesirable byproducts in the
reaction medium. Therefore, oxalic acid is a promising catalyst for
the hydrolysis of cellulose for both the production of sugars and for
pretreatment of the fibers prior to other processes (e.g., enzymatic
hydrolyses).

Acknowledgements

The authors gratefully acknowledge CAPES (Coordination for the


Improvement of Higher Level -or Education- Personnel, Brazil) for
a fellowship for T.M.L. and FAPESP (The State of São Paulo Research
Foundation, Brazil) for financial support. E.F. is grateful to CNPq
(National Research Council, Brazil) for a research productivity fel-
Fig. 7. Arrhenius plot for the acid hydrolysis of sisal pulp (80–100 ◦ C, 4.6 mol L−1 lowship and financial support.
oxalic acid).

References
of sugarcane bagasse using low concentrations of H2 SO4 (2%) at Abushammala, H., Hashaikeh, R., 2011. Enzymatic hydrolysis of cellulose and the
122 ◦ C. use of TiO2 nanoparticles to open up the cellulose structure. Biomass
The activation energy was calculated from the general Arrhe- Bioenergy 35, 3970–3975.
Aguilar, R., Ramírez, J.A., Garrote, G., Vázquez, M., 2002. Kinetic study of the acid
nius law (Eq. (5)), where w is a pre-exponential factor (same hydrolysis of sugar cane bagasse. J. Food Eng. 55, 309–318.
unit as k1 ), Ea is the activation energy, R is the gas constant Arantes, V., Saddler, J.N., 2010. Access to cellulose limits the efficiency of enzymatic
(8.3143 × 10−3 kJ mol−1 K−1 ), and T is the absolute temperature. hydrolysis: the role of amorphogenesis. Biotechnol. Biofuels 3, 1–11.
Ass, B.A.P., Belgacem, M.N., Frollini, E., 2006. Mercerized linters cellulose:
characterization and acetylation in N,N-dimethylacetamide/lithium chloride.
−Ea
lnk1 = + lnw (5) Carbohydr. Polym. 63, 19–29.
RT Balcu, I., Macarie, C.A., Segneanu, A., Pop, R.O., 2011. In: Dr. Shahid Shaukat (Ed.),
Combined Microwave-Acid Pretreatment of the Biomass, Progress in Biomass
Fig. 7 is generated from the logarithm of k1 as a function of the and Bioenergy Production. InTech, pp. 953–978, and references therein
inverse of the absolute temperature to obtain Ea . http://www.intechopen.com/books/progressin-biomass-and-bioenergy-
production/combined-microwave-acid-pretreatment-of-the-biomass
From Fig. 7, a value of 185.9 kJ mol−1 was obtained for the activa- Belgacem, M.N., Gandini, A., 2008. Monomers. In: Polymers and Composites from
tion energy for the hydrolysis of sisal pulp under these conditions. Renewable Resources. Elsevier, London, pp. 560.
The calculated value is higher than the one calculated in a Browning, B.L., 1967. Methods of Wood Chemistry. New York, Interscience.
Budavari, S., 1996. The MERCK index: An Encyclopedia of Chemicals, Drugs and
previous study (60.4 kJ mol−1 ) for the hydrolysis of unmercerized Biological, 12th ed. Whitehouse Station, Merck.
sisal pulp using a sulfuric acid (30%) catalyst at 60–100 ◦ C with Buschle-Diller, G., Zeronian, S.H., 1992. Enhancing the reactivity and strength of
data collected over 6 h (de Paula et al., 2012), and with a higher cotton fibres. J. Appl. Polym. Sci. 45, 967–979.
Bustos, G., Ramirez, J.A., Garrote, G., Vazquez, M., 2003. Modeling of the hydrolysis
concentration of protons in the medium compared to oxalic acid.
of sugarcane bagasse with hydrochloric acid. Appl. Biochem. Biotechnol. 104,
Maleic acid, which is a dicarboxylic acid, has a pKa1 and pKa2 51–68.
(20 ◦ C) of 1.9 and 4.4, respectively, which are relatively close to Ciolacu, D., Pitol-Filho, L., Ciolacu, F., 2012. Studies concerning the accessibility of
different allomorphic forms of cellulose. Cellulose 19, 55–68.
those of oxalic acid. The activation energy for the hydrolysis of
Chen, H., Qiu, W., 2010. Key technologies for bioethanol production from
cellobiose with maleic acid catalysis was reported to be approxi- lignocellulose. Biotechnol. Adv. 28, 556–562.
mately 115 kJ mol−1 (Mosier et al., 2002), which is lower than that Cho, D.H., Shin, S.J., Bae, Y., Park, C., Kim, Y.H., 2011. Ethanol production from acid
determined in this study. However, it should be noted that for hydrolysates based on the construction and demolition wood waste using
Pichia stipites. Bioresour. Technol. 102, 4439–4443.
maleic acid, the reaction occurred under homogeneous conditions Cole, D.E., 2011. Issues facing the auto industry: alternative fuels Technologies, and
because cellobiose is soluble in water. Therefore, issues related Policies. ACP Meeting Eagle Crest Conference Center.
172 T.M. Lacerda et al. / Industrial Crops and Products 71 (2015) 163–172

Da Silva, C.G., Oliveira, F., Ramires, E.C., Castellan, A., Frollini, E., 2012. Composites Mosier, N.S., Sarikaya, A., Ladisch, C.M., Ladisch, M.R., 2001. Characterization of
from a forest biorefinery by-product and agrofibers: lignosulfonate-phenolic dicarboxylic acids for cellulose hydrolysis. Biotechnol. Prog. 17, 474–480.
type matrices reinforced with sisal fibers. Tappi J. 11, 41–49. Okano, T., Sarko, A., 1985. Mercerization of cellulose: II. Alkali–cellulose
Dagnino, E.P., Chamorro, E.R., Romano, S.D., Felissia, F.E., Area, M.C., 2013. intermediates and a possible mercerization mechanism. J. Appl. Polym. Sci. 30,
Optimization of the acid pretreatment of rice hulls to obtain fermentable 325–332.
sugars for bioethanol production. Ind. Crops Prod. 42, 363–368. Pouget, J.P., Józefowicz, M.E., Epstein, J.A., Tang, X., Macdiarmid, A.G., 1991. X-ray
de Paula, M.P., Lacerda, T.M., Zambon, M.D., Frollini, E., 2012. Adding value to the structure of polyaniline. Macromolecules 24, 779–789.
Brazilian sisal: acid hydrolysis of its pulp seeking production of sugars and Ramires, E.C., MegiattoJr, J.D., Gardrat, C., Castellan, A., Frollini, E., 2010. Biobased
materials. Cellulose 19, 975–992. composites from glyoxal-phenolic resins and sisal fibers. Bioresour. Technol.
Dinand, E., Vignon, M., Chanzy, H., Heux, L., 2002. Mercerization of primary wall 101, 1998–2006.
cellulose and its implication for the conversion of cellulose I → cellulose II. Rawat, R., Kumbhar, B.K., Tewari, L., 2013. Optimization of alkali pretreatment for
Cellulose 9, 7–18. bioconversion of poplar (Populusdeltoides) biomass into fermentable sugars
Frollini, E., Bartolucci, N., Sisti, L., Celli, A., 2013. Poly(butylene succinate) using response surface methodology. Ind. Crops Prod. 44, 220–226.
reinforced with different lignocellulosic fibers. Ind. Crops Prod. 45, 160–169. Rocha, G.J.M., Gonçalves, A.R., Oliveira, B.R., Olivares, E.G., Rossell, C.E.V., 2012.
Gámez, S., Ramírez, J.A., Garrote, G., Vázquez, M., 2004. Manufacture of Steam explosion pretreatment reproduction and alkaline delignification
fermentable sugar solutions from sugar cane bagasse hydrolyzed with reactions performed on a pilot scale with sugarcane bagasse for bioethanol
phosphoric acid at atmospheric pressure. J. Agric. Food Chem. 52, 4172–4177. production. Ind. Crops Prod. 35, 274–279.
Gürü, M., Bilgesü, A.Y., Pamuk, V., 2001. Production of oxalic acid from sugar beet Rocha, G.J.M., Nascimento, V.M., Silva, V.F.N.D., Corso, D.L.S., Gonçalves, A.R., 2014.
molasses by formed nitrogen oxides. Bioresour. Technol. 77, 81–86. Contributing to the environmental sustainability of the second generation
Hernández-Salas, J.M., Villa-Ramírez, M.S., Veloz-Rendón, J.S., Rivera-Hernández, ethanol production: delignification of sugarcane bagasse with sodium
K.N., González-César, R.A., Plascencia-Espinosa, M.A., Trejo-Estrada, S.R., 2009. hydroxide recycling. Ind. Crops Prod. 59, 63–68.
Comparative hydrolysis and fermentation of sugarcane and agave bagasse. Rodrigues, F.A., Guirardello, R., 2008. Evaluation of a sugarcane bagasse acid
Bioresour. Technol. 100, 1238–1245. hydrolysis technology. Chem. Eng. Technol. 31, 883–892.
Hoeger, I.C., Nair, S.S., Ragauskas, A.J., Deng, Y., Rojas, O.J., Zhu, J.Y., 2013. Saeman, J.F., Bubl, J.L., Harris, E.E., 1945. Quantitative saccharification of wood and
Mechanical deconstruction of lignocellulose cell walls and their enzymatic cellulose. J. Ind. Eng. Chem. 17, 35–37.
saccharification. Cellulose 20, 807–818. Sathitsuksanoh, N., Zhu, Z., Zhang, Y.-H.P., 2012. Cellulose solvent-based
Hong, B., Xue, G., Guo, X., Weng, L., 2013. Kinetic study of oxalic acid pretreatment pretreatment for corn stover and avicel: concentrated phosphoric acid versus
of moso bamboo for textile fiber. Cellulose 20, 645–653. ionic liquid [BMIM]Cl. Cellulose 19, 1161–1172.
Hongsiri, W., Danon, B., de Jong, W., 2015. The effects of combined catalysis of Shields, S., Boopathy, R., 2011. Ethanol production from lignocellulosic biomass of
oxalic acid and seawater on the kinetics of xylose and arabinose dehydration energy cane. Int. Biodeterior. Biodegrad. 65, 142–146.
to furfural. Int. J. Energy Environ. Eng. 6, 21–30. Sivers, M.V., Zacchi, G., 1995. Atechno-economical comparison of three processes
Koo, B.-W., Min, B.-C., Gwak, K.-S., Lee, S.-M., Choi, J.-W., Yeo, H., Choi, I.-G., 2012. for the production of ethanol from pine. Bioresour. Technol. 51, 43–52.
Structural changes in lignin during organosolv pretreatment of Liriodendron Sjöström, E., 1993. Wood Chemistry: Fundamentals and Applications, 2nd edition.
tulipifera and the effect on enzymatic hydrolysis. Biomass Bioenergy 42, Academic Press, San Diego, California, USA.
24–32. Sullivan, J.M., Willard, J.W., White, D.L., Kim, Y.K., 1983. Production of oxalic acid
Lacerda, T.M., de Paula, M.P., Zambon, M.D., Frollini, E., 2012. Saccharification of via the nitric acid oxidation of hardwood (red oak) sawdust. Ind. Eng. Chem.
Brazilian sisal pulp: evaluating the impact of mercerization on non-hydrolyzed Prod. Res. Dev. 22, 699–709.
pulp and hydrolysis products. Cellulose 19, 351–362. Sun, S., Li, M., Yuan, T., Xu, F., Sun, R., 2013. Effect of ionic liquid/organic solvent
Lacerda, T.M., Zambon, M.D., Frollini, E., 2013. Effect of acid concentration and pulp pretreatment on the enzymatic hydrolysis of corncob for bioethanol
properties on hydrolysis reactions of mercerized sisal. Carbohydr. Polym. 93, production. Part 1: structural characterization of the lignins. Ind. Crops Prod.
347–356. 43, 570–577.
Lee, J.W., Jeffries, T.W., 2013. Efficiencies of acid catalysts in the hydrolysis of Sun, Y., Lin, L., Deng, H., Peng, H., Li, J., Sun, R., Liu, S., 2008. Hydrolysis of bamboo
lignocellulosic biomass over a range of combined severity factors. Bioresour. fiber cellulose in formic acid. Front. For. China 3, 480–486.
Technol. 102, 5884–5890. Sun, R., Lawther, J.M., Banks, W.B., 1995. Influence of alkaline pre-treatments on
Lee, J.W., Houtman, C.J., Kim, H.Y., Choi, I.G., Jeffries, T.W., 2011. Scale-up study of the cell wall components of wheat straw. Ind. Crops Prod. 4, 127–145.
oxalic acid pretreatment of agricultural lignocellulosic biomass for the Taherzadeh, M.J., Karimi, K., 2007. Acid-based hydrolysis processes for ethanol
production of bioethanol. Bioresour. Technol. 102, 7451–7456. from lignocellulosic materials: a review. Bioresources 2, 472–499.
Lenihan, P., Orozco, A., O’Neill, E., Ahmad, M.N.M., Rooney, D.W., Walker, G.M., Tásic, M.B., Konstantinovic, B.V., Lazic, M.L., Veljkovic, V.B., 2009. The acid
2010. Dilute acid hydrolysis of lignocellulosic biomass. Chem. Eng. J. 156, hydrolysis of potato tuber mash in bioethanol production. Biochem. Eng. J. 43,
395–403. 208–211.
Liu, Z.-H., Qin, L., Pang, F., Jin, M.-J., Li, B.-Z., Kang, Y., Dale, B.E., Yuan, Y.-J., 2013. Tian, J., Wang, J., Zhao, S., Jiang, C., Zhang, X., Wang, X., 2010. Hydrolysis of
Effects of biomass particle size on steam explosion pretreatment performance cellulose by the heteropoly acid H3 PW12 O40 . Cellulose 17, 587–594.
for improving the enzyme digestibility of corn stover. Ind. Crops Prod. 44, Tomme, P., Warren, R.A.J., Gikes, N.R., 1995. Cellulose hydrolysis by bacteria and
176–184. fungi. Adv. Microb. Physiol. 37, 1–81.
Lu, Y., Mosier, N.S., 2008. Kinetic modeling analysis of maleic acid-catalyzed vom Stein, T., Grande, P.M., Kayser, H., Sibilla, F., Walter Leitner, W., Domınguez de
hemicellulose hydrolysis in corner stover. Biotechnol. Bioenergy 101, Marıa, P., 2011. From biomass to feedstock: one-step fractionation of
1170–1181. lignocellulose components by the selective organic acid-catalyzed
Ma, Y., Ji, W., Zhu, X., Tian, L., Wan, X., 2012. Effect of extremely low AlCl3 on depolymerization of hemicellulose in a biphasic system. Green Chem. 13,
hydrolysis of cellulose in high temperature liquid water. Biomass. Bioenerg. 39, 1772–1777.
106–111, and references therein. Wang, C., Zhou, F., Yang, Z., Wang, W., Yu, F., Wu, Y., Chi, R., 2012. Hydrolysis of
Martín, M., Grossmann, I.E., 2012. Energy optimization of bioethanol production cellulose into reducing sugar via hot-compressed ethanol/water mixture.
via hydrolysis of switchgrass. AIChE J. 58, 1538–1549. Biomass Bioenergy 42, 143–150.
Martin-Sampedro, R., Revilla, E., Villar, J.C., Eugenio, M.E., 2014. Enhancement of Werle, L.B., Garcia, J.C., Kuhn, R.C., Schwa, M., Foletto, E.L., Cancelier, A., Jahn, S.L.,
enzymatic saccharification of Eucalyptus globulus: steam explosion versus Mazutti, M.A., 2013. Ultrasound-assisted acid hydrolysis of palm leaves
steam treatment. Bioresour. Technol. 167, 186–191. (Roystoneaoleracea) for production of fermentable sugars. Ind. Crops Prod. 45,
Marzialetti, T., Miller, S.J., Jones, C.W., Agrawal, P.K., 2011. Switchgrass 128132.
pretreatment and hydrolysis using low concentrations of formic acid. J. Chem. Xiang, Q., Lee, Y.Y., Torget, R.W., 2004. Kinetics of glucose decomposition during
Technol. Biotechnol. 86, 706–713. dilute-acid hydrolysis of lignocellulosic biomass. Appl. Biochem. Biotechnol.
Megiatto Jr, J.D., Hoareau, W., Gardrat, C., Frollini, E., Castellan, A., 2007. Sisal fibers: 113-116, 1127–1138.
surface chemical modification using reagent obtained from a renewable Xu, J., Chen, H., Kádár, Z., Thomsen, A.B., Schmidt, J.E., Peng, H., 2011. Optimization
source; characterization of hemicellulose and lignin as model study. J. Agric. of microwave pretreatment on wheat straw for ethanol production. Biomass
Food Chem. 55, 8576–8584. Bioenergy 35, 3859–3864.
Morgado, D.L., Frollini, E., Castellan, A., Rosa, D.S., Coma, V., 2011. Biobased films Zhang, S., Keshwani, D.R., Xu, Y., Hanna, M.A., 2012. Alkali combined extrusion
prepared from NaOH/thiourea aqueous solution of chitosan and linter pretreatment of corn stover to enhance enzyme saccharification. Ind. Crops
cellulose. Cellulose 18, 699–712. Prod. 37, 352–357.
Mosier, N.S., Ladisch, C.M., Ladisch, M.R., 2002. Characterization of acid catalytic Zhang, C., Wang, H., Liu, F., Wang, L., He, H., 2013. Magnetic core–shell
domains for cellulose hydrolysis and glucose degradation. Biotechnol. Bioeng. Fe3 O4 @C-SO3 H nanoparticle catalyst for hydrolysis of cellulose. Cellulose 20,
79, 610–618. 127–134, and references therein.

You might also like