Lignin - Kelas B

You might also like

Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 7

Industrial Crops and Products 45 (2013) 296–302

Contents lists available at SciVerse ScienceDirect

Industrial Crops and Products


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

Isolation and characterization of lignin from Moroccan sugar cane bagasse:


Production of lignin–phenol-formaldehyde wood adhesive
a,∗ b b c
Amine Moubarik , Nabil Grimi , Nadia Boussetta , Antonio Pizzi
a MAScIR-NANOTECH, ENSET, Avenue de l’Armée Royale, Madinat El Irfane, 10100 Rabat, Morocco
b Université de Technologie de Compiègne, Unité Transformations Intégrées de la Matière Renouvelable (TIMR, EA4297), Centre de Recherche de Royallieu,
B.P. 20529-60205, Compiègne Cedex, France
c
ENSTIB, Université de Nancy 1, Epinal, France

article info abstract

Article history: lignin-based materials were isolated from Moroccan sugar cane bagasse after alkaline delignification. Sugar cane bagasse was
Received 29 September 2012 ◦ ◦
subjected to hot water (70 C) and alkaline aqueous solutions (15% of sodium hydroxide (NaOH), 98 C) treatments. The
Received in revised form
dissolved lignin macromolecules were separated and purified. The isolated solid was then characterized by different
18 December 2012 Accepted 22 1 13
December 2012 complementary analysis (FT-IR; H, C NMR; GPC and TGA). In the present work, the possibility of preparing wood
adhesives from bagasse lignin has been explored. The results showed that the delignification with 15% NaOH resulted in
yields of cellulose and lignin of 42 ± 2.2% and 13 ± 1.5%, respectively. The extracted lignin scaffolds exhibits high reactivity
Keywords: ◦
Sugar cane bagasse due to the high content of hydroxyl group. Their higher molecular weight (2781 g/mol) and good thermal stability (180 C)
Lignin make them excellent candidates for partial substitution of phenol formaldehyde (PF) resin. A resin formulation in which up to
Wood adhesive 30% of PF can be substituted by bagasse lignin gave good results and was employed for the elaboration of plywood panels
Plywood which passed relevant international standard specifications for interior-grade panels.

© 2013 Elsevier B.V. All rights reserved.

1. Introduction Thangavelu, 2010; Ladeira et al., 2010; Sun and Cheng, 2002). Cellu-lose is a
polydisperse linear homopolymer consisting of regio- and enantioselective -
One of the most challenging topics in material science today is to convert 1,4-glycosidic linked d-glucose units. The poly-mer contains three reactive
biomass-derived waste and feedstock to highly added value-materials. For hydroxyl groups at the C-2, C-3 and C-6 atoms (Heinze and Liebert, 2001;
instance, sugar cane bagasse is a fibrous residue of sugar cane stalks left over Sun et al., 2004). Hemicel-luloses are composed of xylan, mannan, b-glucan,
after crushing and extraction process of the juice from sugar cane. About 54 and xyloglucan polysaccharides (Vibe Scheller and Ulvskov, 2010). Lignin is
million dry tons of bagasse are produced annually throughout the world a complex amorphous polymer composed of phenylpropanoid units consisting
(Mulinari et al., 2009). Most of the bagasse weight is in the form of the so- primarily of coniferyl, sinapyl, and p-coumaryl alco-hols (Hatfield and
called fiber and rind particles, which have high length-width ratios (lengths Vermerris, 2001; Sugimoto et al., 2002). The use of lignin-based materials in
up to a few cm) and correspond mostly to stalk fibro vascular bundles. For the various applications requires a proper investigation of its physico-chemical
sugar industries, this waste is mainly converted into energy through and thermal character-istics to understand the chemical and physical nature
combustion (Leibbrandt et al., 2011), applications in pulping (Goncalves et and thermal behaviour of this bio-polymer. To obtain different products from
al., 2005), acti-vated carbon production (Devnarain, 2003; Qureshi et al., sugar cane bagasse, it is necessary to submit the biomass to sep-aration
2008), cellulosic ethanol production (Carrier et al., 2011) and gasification process of its constituents. However, it is a challenge to isolate original lignin
(Mamphweli and Meyer, 2010). from sugar cane bagasse. A variety of meth-ods like organosolv fractionation
(Pye and Lora, 1991; Balogh et al., 1992), steam explosion (Rocha et al.,
2012; Ibrahim et al., 2010; Glaser and Wright, 1998) and enzymatic
Sugar cane bagasse consists of approximately 50% cellulose, 25% hydrolysis (Shevchenko et al., 1999) have been developed in an attempt to
hemicellulose and 25% lignin (Pandey et al., 2000; Ezhumalai and isolate and iden-tify cellulose and lignin. One of the most suitable methods for
the extraction and separation of lignin is based on the alkaline deligni-fication
(Ibrahim et al., 2010; Sun et al., 2004; Fernandez-Bolanos et al., 1999;
∗ Käuper, 2004; Mousavioun and Doherty, 2010).
Corresponding author at: NANOTECH-MAScIR (Moroccan Foundation for Advanced
Science, Innovation and Research), ENSET, Avenue de l’Armée Royale, Madinat El Irfane,
10100 Rabat, Morocco. Tel.: +212 6 63 52 81 58.
E-mail address: amine.moubarik@yahoo.fr (A. Moubarik).

0926-6690/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.indcrop.2012.12.040
A. Moubarik et al. / Industrial Crops and Products 45 (2013) 296–302 297

Phenol formaldehyde resins are generally synthesized using petro- acetyl groups. Acetylation was performed using Vázquez et al.
chemicals such as phenol and formaldehyde under alkaline catalysts. PF (1999) method.
resins provide high strength and are extremely resis-tant to moisture which
prevent delaminating and enable excellent thermal stability and low initial 2.3. Lignin characterization
viscosity (Pizzi, 1993). There-fore, reduction of cost and substitution of
petroleum-based raw materials are the most important direction of FT-IR measurements were performed in an ABB Bomem FTLA 2000-102
development of PF adhesives. Lignin has been incorporated into wood instrument by direct transmittance using KBr pellet tech-nique. Each spectrum
adhesives due to its similar structure to PF resins (Wang et al., 2009; Turunen was recorded over 20 scans, in the range from 4000 to 200 cm
−1
with a
et al., 2003; Tejado et al., 2007; Kazayawoko et al., 1992). The suitabil-ity of −1
resolution of 4 cm . Background spectra were collected before every
lignin-based materials for manufacturing of lignin–PF resins depends on the
sampling. KBr was previously oven-dried to avoid interferences due to the
polysaccharide composition and total hydroxyl content (phenolic hydroxyl
presence of water. The characteristic bands of lignin were assigned according
and aliphatic hydroxyl). The low con-tent of polysaccharide was researched to the literature.
for the strength and water resistance of the resins, because the presence of
polysaccharide decreased the reactivity of the lignin (Pizzi and Mittal, 1994). 1 13
H and C NMR spectra of extracted lignin samples were recorded on a
In the copolymerization of lignin–PF resins, all of the hydroxyl groups were
Bruker Avance 500 MHz spectrometer from 80 mg of sample dissolved in
important to activate the lignin. DMSO (1.0 mL). Each spectrum was recorded with 32,768 data points, 5.2 s
◦ 13
pulse and pulse delay of 1.547 s (relaxation delay 2.5 s; 90 pulse). The C

In this contribution, lignin fractions isolated from Moroccan sugar cane NMR spectra were recorded at 25 C after 30,000 scans.
bagasse was studied using a combination of several tech-niques TGA, FT-IR,
1 13 Gel permeation chromatography provides a rapid way to obtain
H, C NMR and GPC. The aim of this study was to get a better information on the molecular weight of polymers. Sam-ples have been
understanding on the chemical and thermal prop-erties of the extracted lignin examined through THF-eluted 350A HT-GPC using a MALVERN instrument
sample and to assess their suitability for partial incorporation into phenol (TX, USA) equipped with three styrene–divinylbenzene copolymer gel
formaldehyde (PF) resin. columns of 50, 500, and
´˚
from Polymer Laboratories. The columns were calibrated 104 A
2. Experimental methods
using polystyrene standards in the 92–66.000 g/mol range. The molecular
2.1. Biological material weight and molecular number were compared to polystyrene standard; the
presented results in this work are thus relative values. The flow rate of THF
was 1 ml/min and the samples were dissolved in THF at a concentration of 1
Sugar cane bagasse was obtained from a local sugar factory (Doukkala, ◦
Morocco). The bagasse was stored indoors during the experiments. The mg/ml and stored for 24 h at 5 C to avoid variations in molecular weight.
bagasse was air-dried for three days at ambient temperature (up to 8–10%
equilibrium moisture content) and then cut into small pieces (1–2 cm). The Thermogravimetric analysis (TGA) was used to determine the thermal
stability and degradation of the lignin samples using a TGA Q50
cut bagasse was ground to pass a 1.0 mm size screen. All chemicals used
thermogravimetric apparatus. Ten milligrams of each cured sample were
were of analytical or reagent grade (Sigma–Aldrich, France).
placed on a balance located in the furnace and heat was applied over the
◦ ◦
temperature range from room temperature to 1000 C at a heating rate of 5
2.2. Isolation of lignin C/min in air. Mass losses vs. temperature thermograms were obtained
showing the different decomposition processes. Three replicates were used for
optimal adhesive mix.
The procedure for isolation of lignin by delignification with aqueous
solution of sodium hydroxide is illustrated in Fig. 1. Twenty grams (oven-dry

matter) of sugarcane bagasse were treated with a hot water at 70 C for 2 h.
The solid-to-liquid ratio used was 1:10. The treatments were carried out in a 2.4. Preparation of resin formulation
steel reactor with a Parr 4836 temperature controller (Parr Instrument
Company, Moline, IL). At the end of the reaction the pre-treated sugarcane A phenol-formaldehyde resol with a solids content of 46% and a viscosity
◦ of about 450 cp was prepared using a 2.2:1 formaldehyde: phenol ratio and
bagasse was cooled (25 C), washed with water (S:L ratio = 1:10; w/w) and
7.3% (w/w) of NaOH. The resols were prepared in a two liters glass reactor
centrifuged (2000 rpm, t = 10 min) to separate the solubilized hemi- with mechanic stirring and tempera-ture control. The necessary amount of the
celluloses. reactive according to the established formulation was fed into the reactor.

When the operat-ing temperature was reached (90 C), the extension of
Lignin was then extracted from the pretreated sugar cane bagasse using

alkaline treatment. Sugar cane bagasse was placed in a reactor with an reaction was monitored, measuring resol viscosity at 25 C. The lignin–PF
aqueous alkaline (15% NaOH (m/v)) solution. The solid-to-liquid ratio was adhe-sives were prepared by copolymerisation of lignin (with variable
1:10. The suspension was maintained under agitation (250 rpm) for 90 min at amounts) at room temperature.

98 C. At the end of the reaction the delignified material was filtered to
obtain the black liquor without any fibrous materials. Sulfuric acid (5 N H 2
2.5. Plywood manufacture and testing
SO4 ) was added until reaching pH 2 in order to precipitate the acidified
lignin, which was then collected and washed by centrifugation and air-dried.
The lignin powder, with rather low amount of carbohydrate, was used directly Five ply laboratory plywood panels of dimension 250 mm × 250
in the preparation of lignin–PF adhesive. No addi-tional purification steps mm × 10 mm were prepared from 2 mm thick Maritime pine (Pinus pinaster)
were performed. All the yields of lignins represent the mean of at least 2
veneers of 4% moisture content at a glue mix spread of 225 g/m single glue
triplicate analysis. line. Plywood bonded with bagasse lignin–PF resin has been assembled and

hot pressed under 12 bar at 160 C for 6 min. The fixed bonding conditions of
Lignin samples were subjected to acetylation in order to enhance their
1 ◦
solubility in organic solvent in gel permeation chro-matography (GPC) and 160 C pressing tempera-ture and 12 bar were selected to reproduce the
13 industrial conditions used to bond plywood panels. The longer pressing time
H, C NMR analysis. During the acetylation process, all the hydroxyl
of 6 min was used to assure full reaction.
functional groups are substituted by new
298 A. Moubarik et al. / Industrial Crops and Products 45 (2013) 296–302

Sugar cane
bagasse

Extracted with hot water at 70°C during


2 hours (solid:liquid ratio; 1:10, w/w).

Separated
sample

Extracted with 15% NaOH (w/v)


aqueous solution at 98 °C for 90 min
(solid:liquid ratio; 1:10, w/w)

Residue : Filtrate
cellulose

Neutralized with 5N H2SO4 to pH 2

Acidified
lignin

Washing with water and air-drying

Lignin

Fig. 1. Scheme for extractions of original lignins from Moroccan sugar cane bagasse.

Mechanical properties commonly taken into consideration in the general calculated. The analysis of variance (ANOVA) was applied for the analysis of
usage areas of plywood panels were investigated. Dry tensile strength and the results of Table 5. For each analysis, significance level of 5% was
modulus of elasticity values of plywood panels were determined according to assumed. All statistical analyses were carried out using the software
EN 314 (1993) and EN 310 (1993), respectively. Fifty specimens were used Statgraphics Plus 5.1 (Stat-point Technologies, Inc.).
for each test method, and the results obtained are shown in Table 5.

3. Results and discussion


2.6. Formaldehyde emission by desiccator method
3.1. Yield of isolated lignins
The formaldehyde emissions from the plywood were deter-mined
according to the European Norm (ISO/CD 12460-4) using a glass desiccator. Alkaline treatment is usually an effective method to extract lignin from
The 24-h desiccator method uses a common glass desiccator with a volume of agricultural residues. Sodium hydroxide, one of the most common alkaline
10 L. Eight test pieces, with dimen-sions of 150 mm × 50 mm × 10 mm, reagents, has been applied to treat a vari-ety of agricultural residues (Kumar
which were cut from plywood, are positioned in the desiccator. The and Wyman, 2009; Ong et al., 2010). In this study, alkaline lignins were

formaldehyde released from the test pieces at 23 ± 2 C and 50 ± 10% isolated with 15% NaOH at 98 ◦ C for 90 min. Yields of lignin and cellulose
relative humidity, during 24 h is absorbed in a Petri dish filled with a 30 ml of were calculated on a dry weight basis. The obtained yields of lignin and
distilled water. The released amount was then determined photometrically. cellulose are respectively 13% and 42%. These values are in perfect
Thirty replicates were used for each adhesive. consistency with the literature, which reports cellulose and lignin percentages
of 50% and 15% (Pandey et al., 2000; Ezhumalai and Thangavelu, 2010;
with hot
Ladeira et al., 2010), respectively. The two-stage treatments
2.7. Statistical analysis water and 15% NaOH, totally released 86% of original lignin
Each experiment was repeated, at least, 10 times. The tests experiments
and 84% of original cellulose from sugar cane bagasse. This
concerning the plywood characterization were repeated 50 times. Means and sequen-tial treatment enabled the isolation of original lignin
standard deviations of data were from sugar can bagasse.
A. Moubarik et al. / Industrial Crops and Products 45 (2013) 296–302 299
Table 1

−1 Vibration Assignment
Band (cm )
3400–3405 st O H Phenolic OH and aliphatic OH
2960–2925 st C H CH3 and CH2 groups
2850–2840 st C H OCH3
∼1600 st C C Aromatic skeleton
1460 C H deformation Asymmetric in CH3 and CH2
1425 st C C Aromatic skeleton
1365 ı O H Phenolic OH
ip

∼1220 st C O(H) + C O(Ar) Phenolic OH and ether


1125 ıip Ar C H Guaiacyl (G)
1115 ıip Ar C H Syringyl (S)
1030 st C O(H) + C O(C) First order aliphatic OH and ether
855 ıop Ar C H Guaiacyl (G)

Fig. 2. FTIR spectra of the bagasse lignin. st: stretching vibration.


ıip : in-plane deformation vibration.
ıop : out-of-plane deformation vibration.

3.2. Lignin characterization Further information on the chemical structure of the isolated lignin
13
samples was obtained by C NMR. The spectra of lignin are shown in Fig.
The FT-IR spectra of extracted lignins are illustrated in Fig. 2. FT-IR 3. In order to illustrate the distinct attributions, the corresponding assignments
spectra reflects the chemical structure as well as the purity of lignins. The identified for the lignin are listed in Table 2 according to the previous
corresponding assignments and bands for lignins are presented in Table 1. literatures (Pan et al., 1994; Peter et al., 1996). The major peaks of the
The presence of syringyl (S) and guaia-cyl (G) bands for lignins indicate that syringyl rings appear at 102 ppm (C2 and C6 ). Furthermore, the characteristic
the lignin extracted from sugar cane bagasse is more similar to wood lignin peaks of the side alkyl chains are shown at 76–73 ppm (C OH in b-O-4 linked
than annual plant lignin, which is normally HGS lignin. This similarity is due side chain) and 63–60 ppm (C with C O and C O in p-coumaryl ester). In
−1
to the absence of the band of hydroxyl phenyl propane (H) at 1166 cm in addition, the signal corresponding to the methoxyl group is observed at around
the spectra (Zhao et al., 2009). The presence of guaiacyl-type (G) confirms 56–57 ppm. A very weak resonance of carbo-hydrates between 90 and 100
that bagasse lignins had potential active site for polymer-ization. The ppm indicates low concentration of residual sugars in lignin (Yuan et al.,
presence of G-type unit revealed that bagasse lignin could react with 2011). This observation agrees with the results of the FTIR spectra, in which
formaldehyde and could be cross linked with formaldehyde in the same way the typical bands of cellulose were not found.
as in the phenol formaldehyde con-densation reaction. The presence of the
−1
signal band at ∼1460 cm , assigned to C H deformation (asymmetric) in
1
methyl, methylene and methoxyl groups, confirms that lignin aromatic The H NMR spectra of bagasse lignin are shown in Fig. 4. Table 3
structures did not change dramatically during the soda extraction procedure shows the hydrogen signal integrations subdivided into differ-ent structural
(Nadji et al., 2009). regions (Goncalves et al., 2000; Guerra et al., 2004; Hiltunen et al., 2006). In
1
the H NMR spectrum, the two signals at 2.50 and 3.46 ppm arise from
DMSO-d6 and HDO, respectively. The

13
Fig. 3. C NMR spectrum of the bagasse lignin.
300 A. Moubarik et al. / Industrial Crops and Products 45 (2013) 296–302

Table 2 Table 4

Peak (ppm) a Assignment Lignin Lignin


Intensity
acetylated
102 vs C2 , C6 (S)
76–73 vs C OH in b-O-4 linked; side chain Mw (g/mol) 1315 2781
63 s C- in -5, C- in coniferyl alcohol units Mn (g/mol) 699 1365
56 w Methoxyl group Mw /Mn (polydispesity) 1.88 2.03
a vs: very strong; s: strong; w: weak; vw: very weak.

Table 3
of and -O-4 bonds is the predominant process in alkaline medium. Table 4
Peak (ppm) Assignment shows that the polydispersity for bagasse lignins was relatively low, indicating
8.0–6.0 Aromatic H in S and G units that bagasse lignins have a high fraction of low-molecular weight molecules.
6.9 Aromatic H in G Lignins with high fractions of low molecular weight molecules are very
6.6 Aromatic H in S suitable for condensation with PF because they are more reactive than those
3.1–4.2 Methoxyl H with high molecular weight molecules (Pizzi and Mittal, 1994).
2.5–2.2 H in aromatic acetates
2.2–1.9 H in aliphatic acetates
1.5–0.8 Aliphatic H The TGA and DTG curves of lignin obtained by sequential extrac-tion
from bagasse are displayed in Fig. 5. As it can be seen, the decomposition of

bagasse lignins covered wide temperature ranges from 200 to 1000 C. The
broad signal at 5.31 ppm corresponds to H- of -5 structures. The signals at degradation stage can be divided into three stages. The initial weight loss step
4.23 ppm and signals between 4.84 and 5.28 ppm were attributed to H- , H- ◦
occurs at 30–120 C due to the evaporation of absorbed water. The second
and OH in -O-4 , respectively. The signal of H- in -O-4 was overlapped with ◦
stage occurred at around 180–350 C and is ascribed to the degradation of the
that of HDO. The small peak at 1.20 ppm corresponds to methylene groups of
carbohydrate components of the lignins, which were converted to volatile
aliphatic chain.
The molecular weight of lignin samples have been analyzed through THF- gases including CO, CO2 , and CH4 . After that, lignin was degraded over a

eluted GPC. As shown before, lignin samples have been acetylated to enhance wide range of temperature above 350 C. The degradation volatile products
their solubility in such solvent. The weight-average (M w ) and number- derived from lignin were phenolics, alcohols, aldehyde acids with the
average (Mn ) molecular weight, and polydispersity (M w /Mn ) of the lignin formation of gaseous products (CO, CO 2 and CH4 ) (Liu et al., 2008). The
before and after acety-lation were determined (Table 4). It was observed that TGA curve also shows that thermal degradation began to occur only after the
materials have absorbed certain amounts of heat energy. The heat initiates the
the Mw of lignin samples before being acetylated are comparatively lower degradation processes and the breaking down of the struc-ture by causing
than those after acetylation. This observation indicates that the ◦
macromolecules of lignin were not completely dissolved in THF-eluted GPC. molecular chain ruptures. At 800 C, about 20% of non-volatile residue, still
remained in solid form and were not completely burned. This result reveals
Thus, acetylation should be carried out. Soda bagasse lignins have lower M w that bagasse lignins are sta-ble at high temperature, which is attributed to the
and Mn than organosolv and ethanol process lignins (El Mansouri and high degree of branching and formation of highly condensed aromatic
Salvado, 2006). This result corroborates the higher phenolic hydroxyl in these structure
lignins in which the cleavage

Fig. 4. 1
H NMR spectrum of the bagasse lignin.
A. Moubarik et al. / Industrial Crops and Products 45 (2013) 296–302 301


Fig. 5. TGA and DTG curves of the bagasse lignin, ˇ = 5 C/min in air.

for bagasse lignins. The results obtained with thermogravimetric analysis



confirm that in 180 C these lignins are not damaged.

3.3. Lignin valorization

Sugar cane bagasse lignin was selected and used for the elab-oration of
wood adhesive. The dry tensile strength, wood failure, modulus of elasticity
(MOE) and modulus of rupture (MOR) val-ues of plywood glued with resins
made with different weight ratio of lignin/PF (10/90, 30/70 and 50/50) was

tested and compared to a control PF resin (0:100) at 160 C and 6 min press
time. Table 5 reports the results of the effect of lignin substitution level on the
mechanical performance of plywood panels. Up to 30% of lignin, the bond
strength and wood failure appear to be relatively unaf-fected. Furthermore,
the test reveals that in most cases, joint failure was cohesive in the wood
(wood failure) and it was not due to failure either at the interface or at the
adhesive itself. Compared to native PF resins, the addition of lignin up to 30%
improves the MOE and MOR by 14% and 79%, respectively. Statically
Fig. 6. FTIR spectra of PF resin and 30% lignin–PF resins.
significant difference was present between the control and plywood manu-
factured with lignin–PF (30/70) resin. However, as it can be seen from Table
5, resin 50/50 exhibited dramatically low mechanical performances (tensile
strength, wood failure, MOE and MOR), com-pared to the other resins. with lignin–PF (30/70) and commercial PF resins. The reduction of
Finally, plywood panels with lignin–PF (30/70) resin shows slight reduction formaldehyde emission of plywood panels is due to the lignin sub-stitution of
(40%) in formaldehyde emis-sion compared to panels prepared with PF resins and the PF penetration in the wood. Beyond 30% a sharp increase in
commercial PF resins. The ANOVA showed a significant difference between formaldehyde emission is observed. This is generally due to poor PF
panels bonded penetration in the wood.
Fig. 6 shows a representative FTIR spectrum of PF and opti-mal resins
composition (30% lignin–PF). This measurement was

Table 5

Formulations Composition Dry tensile strength Wood failure, MOE (MPa) MOR (MPa) Formaldehyde emission
lignin/PF (w/w) (MPa) mean ± SD mean (%) mean ± SD mean ± SD 2
(mg/m /h) mean ± SD
0% LPF 0/100 a a 3114 ± 109a 44 ± 5a 2.52 ± 0.20a
1.2 ± 0.04 70
(control) b a a b
± 0.10 ± 265 ±4 ± 0.17
10% LPF 10/90 1.5 b 2907 47 2.03
60
30% LPF 30/70 1.8 ± 0.08c 75
c
3578 ± 98b 79 ± 8b 1.51 ± 0.12c
50% LPF 50/50 0.8 ± 0.11d 20
d 1734 ± 231c 32 ± 11a 2.11 ± 0.17ab

a, b, c, d values with the same superscript letters were not significantly different; SD: standard deviation.
MOR: modulus of rupture.
MOE: modulus of elasticity.
302 A. Moubarik et al. / Industrial Crops and Products 45 (2013) 296–302

used to monitor the reaction between PF and bagasse lignin. Spec-tral Ibrahim, M.M., Agblevor, F.A., El-zawawy, W.K., 2010. Isolation and characterization of
−1 cellulose and lignin from steam-exploded lignocellulosic biomass. BioRe-sources 5 (1),
difference between PF and lignin–PF resins was observed at 980 cm , 397–418.
which was assigned to C H stretching vibration of vinyl in the PF. The peaks Käuper, P., 2004. From production to product, Part 1. Solution states of alkaline bagasse lignin
−1 solution. Ind. Crop. Prod. 20, 151–157.
at 1020 cm , were ascribed to the presence of C O stretching vibration of
Kazayawoko, J.S.M., Riedl, B., Poliquin, J., Barri, A.O., Matuana, L.M., 1992. A lignin–
aliphatic C OH, aliphatic C O (Ar) and methylol C OH. The peaks at 1020 phenol-formaldehyde binder for particleboard: part 1. Holzforschung 46, 257–262.
−1
cm of lignin–PF resins were broader than that of the PF resin, which is
Kumar, R., Wyman, C.E., 2009. Effects of cellulase and xylanase enzymes on the
attributed to the pres-ence of other species (not only methylol–OH but also
deconstruction of solids from pretreatment of poplar by leading technologies. Biotechnol.
−1
alphatic–OH in lignin–PF resins). The bond at 1250 cm of PF resin was Prog. 25, 302–314.
bigger than that in the lignin–PF resins, indicating more C O stretching of Ladeira, N.C., Peixoto, V.J., Penha, M.P., Paula Barros, E.B., Ferreira Leite, S.G., 2010.
Optimization of 6-pentyl- -pyrone production by solid state fermentation using sugarcane
phenolic–OH groups in PF resin than in lignin–PF resins.
bagasse as residue. BioResources 5 (4), 2297–2306.
Leibbrandt, N.H., Knoetze, J.H., Görgens, J.F., 2011. Comparing biological and ther-
mochemical processing of sugarcane bagasse: an energy balance perspective. Biomass
4. Conclusion Bioenergy 35 (5), 2117–2126.
Liu, Q., Wang, S.R., Zheng, Y., Luo, Z.Y., Cen, K.F., 2008. Mechanism study of wood lignin
In this work, the composition and physicochemical properties of lignin pyrolysis by using TG–FTIR analysis. J. Anal. Appl. Pyrol. 82, 170–177.
3
isolated from Moroccan sugar cane bagasse by alkali treat-ments were Mamphweli, N.S., Meyer, E.L., 2010. Evaluation of the conversion efficiency of the 180 Nm /h
TM
1 13 Johansson biomass gasifier . Int. J. Energy Environ. 1 (1), 113–120.
investigated. FT-IR spectroscopy and H, C NMR spectrometry reveals
that lignin is mainly composed of G and S units and less free hydroxyls and Mousavioun, P., Doherty, W.O.S., 2010. Chemical and thermal properties of fraction-ated
bagasse soda lignin. Ind. Crop. Prod. 31 (1), 52–58.
highlights the absence of resid-ual sugars in lignin. Molecular weight of Mulinari, D.R., Voorwald, H.J.C., Cioffi, M.O.H., Da Silva, M.L.C.P., Luz, S.M., 2009.
lignin samples has been studied through THF-eluted GPC showing that Preparation and properties of HDPE/sugarcane bagasse cellulose composites obtained for
bagasse lignin has much higher M W and Mn . This behavior is related to its thermokinetic mixer. Carbohydr. Polym. 75, 317–321.
Nadji, H., Diouf, P.N., Benaboura, A., Bedard, Y., Riedl, B., Stevanovic, T., 2009. Com-
higher content in C C linkages. Both structural and thermal properties suggest
parative study of lignins isolated from alfa grass (Stipa tenacissima L.). Bioresour. Technol.
that bagasse lignin will be a better substitute in the syn-thesis of lignin–PF 100, 3585–3592.
resins, as it presents higher amount of activated free ring positions, higher Ong, L.G.A., Chuah, C., Chew, A.L., 2010. Comparison of sodium hydroxide and potassium
MW and higher thermal decomposition temperature. hydroxide followed by heat treatment on rice straw for cel-lulose production under solid
state fermentation. J. Appl. Sci. 10 (21), 2608–2612.

Pan, X.Q., Lachenal, D., Neirinck, V., Robert, D., 1994. Structure and reactivity of spruce
13
Lignin–based wood adhesives prepared with PF, were prepared and tested mechanical pulp lignins: C NMR spectral studies of isolated lignins. J. Wood Chem.
Technol. 14, 483–506.
for application to wood panels. As a proof of concept, bagasse lignin could be Pandey, A., Soccol, C.R., Nigam, P., Soccol, V.T., 2000. Biotechnological poten-tial of agro-
used to replace 30% of the PF resins used to bond plywood panels, without industrial residues, I: sugarcane bagasse. Bioresour. Technol. 74, 69–80.
adversely affecting bond properties. Moreover, the formaldehyde emission
Peter, M.F., Arthur, J.R., Jiang, J.E., 1996. Chemical structure of residual lignin from kraft pulp.
levels obtained from pan-els bonded with lignin–PF (30% lignin–PF) were J. Wood Chem. Technol. 16, 347–365.
lower than those obtained from panels made with control PF. Pizzi, A., 1993. Wood Adhesives Chemistry and Technology, vol. 1. Marcel Dekker, New York.

Pizzi, A., Mittal, K.L. (Eds.), 1994. Handbook of Adhesives Technology. Marcel Dekker, New
York.
References Pye, E.K., Lora, J.H., 1991. The ALCELL process: a proven alternative to kraft pulping.
TAPPI 74 (3), 113–118.
Balogh, D.T., Curvelo, A.A.S., De Groote, R.A., 1992. Solvent effect on organosolv lignin from Qureshi, K., Bhatti, I., Kazi, R., Ansari, A.K., 2008. Physical and chemical analysis of activated
Pinus Caribaea Hondurensis. Holzforschung 46, 343–348. carbon prepared from sugarcane bagasse and use for sugar decolori-sation. Int. J. Chem.
Carrier, M., Hugo, T., Gorgens, J., Knoetze, J.H., 2011. Comparison of slow and vacuum Biol. Eng. 1, 3.
pyrolysis of sugar cane bagasse. J. Anal. Appl. Pyrol. 90, 18–26. Rocha, G.J.M., Gonc¸ alves, A.R., Oliveira, B.R., Olivares, E.G., Rossell, C.E.V., 2012. Steam
Devnarain P.B., 2003. Production of activated carbon from South African sugar-cane bagasse. explosion pretreatment reproduction and alkaline delignification reactions per-formed on a
M.Sc. (Eng.) Thesis. University of KwaZulu-Natal. pilot scale with sugarcane bagasse for bioethanol production. Ind. Crop. Prod. 35 (1), 274–
El Mansouri, N.E., Salvado, J., 2006. Structural characterization of technical lignins for the 279.
production of adhesives: application to lignosulfonate, kraft, soda-anthraquinone, Shevchenko, S.M., Beatson, R.P., Saddler, J.N., 1999. The nature of lignin from steam
organosolv and ethanol process lignins. Ind. Crop. Prod. 24, 8–16. explosion/enzymatic hydrolysis of softwood. Appl. Biochem. Biotechnol. 77–79, 867–876.

EN 310, 1993. Wood-based panels. Determination of modulus of elasticity in bending and of Sugimoto, T., Akiyama, T., Matsumoto, Y., Meshitsuka, G., 2002. The erythro/threo ratio of
bending strength. beta-O-4 structures as an important structural characteristic of lignin – part 2, changes in
EN 314-1/-2, 1993. Plywood-Bond Quality, Part 1. Test Methods. European Commit-tee for erythro/threo (E/T) ratio of beta-O-4 structures during delig-nification reactions.
Standardization, Brussels. Holzforschung 56, 416–421.
Ezhumalai, S., Thangavelu, V., 2010. Kinetic and optimization studies on the biocon-version of Sun, J.X., Sun, X.F., Zhao, H., Sun, R.C., 2004. Isolation and characterization of cellulose from
lignocellulosic material into ethanol. BioResources 5, 1879–1894. sugarcane bagasse. Polym. Degrad. Stab. 84, 331–339.
Fernandez-Bolanos, J., Felizon, B., Heredia, A., Guillén, R., Jiménez, A., 1999. Charac- Sun, Y., Cheng, J., 2002. Hydrolysis of lignocellulosic materials for ethanol produc-tion: a
terization of the lignin obtained by alkaline delignification and of the cellulose residue from review. Bioresour. Technol. 83 (1), 1–11.
stem-exploded olive stones. Bioresour. Technol. 68, 121–132. Tejado, A., Pena, C., Labidi, J., Echevérria, J.M., Mondragon, I., 2007. Physic-chemical
Glaser, W.G., Wright, R.S., 1998. Steam assisted biomass fractionation. II. Fraction-ation characterization of lignins from different sources for use in phenol-formaldehyde resin
behavior of various biomass resources. Biomass Bioenergy 14, 219–235. synthesis. Bioresour. Technol. 98 (8), 1655–1663.
Goncalves, A.R., Ruzene, D.S., Moriya, R.Y., Oliveria, L.R.M., 2005. Pulping of sugarcane
bagasse and straw and biobleaching of the pulps: conditions parameters and recycling of Turunen, M., Alvila, L., Pakkanen, T.T., Rainio, J., 2003. Modification of phenol-formaldehyde
enzymes. In: 59th Appita Conference, Auckland, New Zealand, 16–19 May. resol resins by lignin, starch and urea. J. Appl. Polym. Sci. 88 (2), 582–588.

Goncalves, A.R., Schuchardt, U., Bianchi, M.L., Curvelo, A.A.S., Braz, J., 2000. Piassava Vázquez, G., Freire, S., Bona, C.R., González, J., Antorrena, G., 1999. Structures and
fibers (Attalea Funifera): NMR spectroscopy of their lignin. J. Braz. Chem. Soc. 11, 491– reactivities with formaldehyde of some acetosolv pine lignins. J. Wood Chem. Technol. 19,
494. 357–378.
Guerra, A., Mendonca, R., Ferraz, A., Lu, F., Ralph, J., 2004. Structural characterization of Vibe Scheller, H., Ulvskov, P., 2010. Hemicelluloses. Annu. Rev. Plant Biol. 61, 263–289.
lignin during Pinus Taeda wood treatment with Ceriporiopsis Subvermispora. Appl.
Environ. Microbiol. 70, 4073–4078. Wang, M., Leitch, M., Xu, C., 2009. Synthesis of phenol formaldehyde resol resins using
Hatfield, R., Vermerris, W., 2001. Lignin formation in plants, the dilemma of linkage organosolv pine lignins. Eur. polym. J. 45 (12), 3380–3388.
specificity. Plant Physiol. 126, 1351–1357. Yuan, T.Q., Sun, S., Xu, F., Sun, R.C., 2011. Isolation and physic-chemical char-acterization of
Heinze, T., Liebert, T., 2001. Unconventional methods in cellulose functionalization. lignin from ultrasound irradiated fast-growing poplar wood. BioResources 6 (1), 414–433.
Prog. Polym. Sci. 26, 1689–1762.
Hiltunen, E., Alvila, L., Pakkanen, T.T., 2006. Characterization of Brauns’ lignin from fresh Zhao, X.B., Dai, L., Liu, D., 2009. Characterization and comparison of acetosolv and milox
and vacuum-dried birch (Betula Pendula) wood. Wood. Sci. Technol. 40, 575–584. lignin isolated from crofton weed stem. J. Appl. Polym. Sci. 114, 1295–1302.

You might also like