Vous êtes sur la page 1sur 7

Polymer Degradation and Stability 109 (2014) 33e39

Contents lists available at ScienceDirect

Polymer Degradation and Stability


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

Impact of catalytic oil palm fronds (OPF) pulping on organosolv lignin


properties
M. Hazwan Hussin a, b, Adah Abdul Rahim a, Mohamad Nasir Mohamad Ibrahim a,
Dominique Perrin b, Mehdi Yemloul c, Nicolas Brosse b, *
a

Lignocellulosic Research Group, School of Chemical Sciences, Universiti Sains Malaysia, 11800 Minden, Penang, Malaysia
Laboratoire d'Etude et de Recherche sur le MAteriau Bois (LERMAB), Faculte des Sciences et Techniques, Universite de Lorraine, Bld des Aiguillettes,
F-54500 Vandoeuvre-les-Nancy, France
c
^me, F-13397 Marseille, France
Institut de Sciences Moleculaire de Marseille, Aix-Marseille Universite, Service 512, Campus Scientique de St J
ero
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 19 May 2014
Received in revised form
19 June 2014
Accepted 22 June 2014
Available online 5 July 2014

This article sheds light on the structural characteristic and antioxidant activity of the ethanol organosolv
lignin extracted from oil palm fronds (OPF) via incorporation of 1,8-dihydroxyanthraquinone during the
delignication process. The resulting modied organosolv lignin (DEOL) was studied by 31P NMR, HSQC,
HMBC and GPC. It was proposed that addition of a catalytic amount of 1,8-dihydroxyanthraquinone
during pulping process; (1) enhanced the dissolution of lignin and the delignication rate, (2)
improved the solubility of the resulting modied lignin (DEOL) by reducing its hydrophobicity properties
and (3) improved its antioxidant activity compared to untreated organosolv lignin (EOL) (DEOL: 78% and
EOL: 53% of Oxygen Uptake Inhibition (OUI) respectively). It was shown that antioxidant activity was
closely related to its average molecular weight and phenolic hydroxyl content.
2014 Elsevier Ltd. All rights reserved.

Keywords:
Organosolv lignin
Oil palm fronds
1,8-Dihydroxyanthraquinone
Phenolic eOH
Antioxidant

1. Introduction
The most crucial issues faced by the world today are to ensure
the sustainability of consumption for energy and natural resources.
As the fossil fuel is creating problematic issues (such as global
warming, increase in price and running out), the use of renewable
resources to shift the oil-based economy into bio-based economy
leads to a new discovery as an alternative. With a goal of reducing
net greenhouse gas emission, this marks an important turning
point in effort to promote the use of renewable energy to fulll the
commitments of the Kyoto Protocol [1,2]. Lignocellulosic biomass is
best-suited for energy and chemical applications due to its sufcient availability, inexpensive and environmentally safe. Recent
work in this area has mainly focused on the delignication of
lignocellulosic biomass separating lignin, cellulose and hemicelluloses to be used in both physical and chemical applications.
It has been acknowledged that organosolv delignication allows
a clean fractionation of lignocellulosic feedstocks and the recovery
of high-quality lignins (relatively pure, less condensed than other
industrial lignins, sulfur free, soluble in organic solvent) are of great

* Corresponding author. Tel.: 33 3 83 68 48 62; fax: 33 3 83 68 44 98.


E-mail address: Nicolas.Brosse@lermab.uhp-nancy.fr (N. Brosse).
http://dx.doi.org/10.1016/j.polymdegradstab.2014.06.016
0141-3910/ 2014 Elsevier Ltd. All rights reserved.

interest and are currently a focus of attention [3]. Thus, availability


of such organosolv lignin fractions in large quantities should
stimulate development in new lignin utilizations. However, the
valuable utilization of the lignin produced at the industrial scale
requires a good control of its variability which is a function of the
nature of the raw material and also of the processes used for the
lignin extraction [4].
Lignin can act as a neutralizer or inhibitor in oxidation processes, via stabilizing reactions induced by oxygen radicals and
their respected species due to high content of diverse functional
groups (phenolic and aliphaticeOH, carbonyls, carboxyls, etc.) and
its phenylpropanoid structure. The applicability of lignins from
different sources as potential antioxidants has been also successfully tested [5e7]. Moreover, it was revealed that the extraction
processes of lignin may give major effect on its antioxidant capacity
[8]. Consequently, the antioxidant properties exhibited by lignin
can give broader applications as anti-microbial, anti-aging agents
and corrosion inhibitors.
Nevertheless, the complexity of lignin structures might jeopardize the fate of these organic compounds. In addition, high hydrophobicity of organosolv lignin can limit its capability to be
employed in other possible applications. Therefore, the modulation
of suitable lignin structures (by considering its solubility, molecular
weight, phenolic content) is important so that it can overcome such

34

M.H. Hussin et al. / Polymer Degradation and Stability 109 (2014) 33e39

implications. Obviously, the properties of lignin can be improved by


modifying the structure into a more suitable structure type. Electropolymerization [9] and graft polymerization [10] of lignin are
some effective ways to convert insoluble lignin into a water-soluble
lignin. It was also demonstrated that the presence of carbonium ion
scavengers in combinative treatment (prehydrolysis followed by
delignication) could substantially improve the lignin extraction
yields and its properties through inhibition of recondensation reactions [7,11e13]. In this approach, the lignin deconstruction during
the prehydrolysis treatment increases its extractability on organosolv treatment through the breaking of ligninecarbohydrate
bonds, resulting in smaller lignin fragments. Smaller and compact
structure (low molecular weight, high phenolic eOH content with
better solubility) of lignin is indeed benecial for later usage
especially in antioxidant applications.
Addition of anthraquinone in soda pulping is described to
accelerate wood delignication. Anthraquinone has been shown to
operate in a redox cycle: the reducing end groups of carbohydrates
dissolved in the pulping liquor reduce the anthraquinone to
anthrahydroquinone (Fig. 1). In the soda pulping conditions,
anthrahydroxyquinone then reacts with lignin quinone methides
through an additioneelimination reaction and causes the b-aryl
ether linkage in the lignin molecule to cleave [14,15].
In the previous work, we have successfully revealed the effect of
2-naphthol as an organic scavenger during combinative treatment
of oil palm fronds (OPF) biomass [7]. Even though incorporation of
lignin with 2-naphthol improves the lignin extraction yield and its
antioxidant activity (Oxygen Uptake Index (OUI) %: ~70%), it was
found that the solubility in water (~2.07% dissolution) of this type of
lignin was very low. Higher solubility of lignin is somewhat
important for subsequent applications. Therefore, in order to
improve the properties of lignin, the aim of the present work was to
investigate the impact of the incorporation of another suitable aromatic
moieties
derived
from
anthraquinone
(1,8dihydroxyanthraquinone) during the delignication process on
the lignin physicochemical properties. The inuence of lignin
structure by its (syringyl) S, (guaiacyl) G and (p-hydroxyphenyl) H
basic units, phenolic/aliphatic hydroxyl content and molecular size
were also identied via FTIR, 31P NMR and GPC. Besides, the effect
on
the
lignin
structure
in
the
presence
of
1,8dihydroxyanthraquinone as a carbonium ion scavenger was also
studied with additional information obtained from two dimensional 13Ce1H HSQC, HMBC NMR. The improvement of lignin solubility and antioxidant properties were also investigated via
dissolution test, oxygen uptake inhibition and reducing power
assay.

2. Material and methods


2.1. Material
The oil palm fronds (OPF) were obtained from Valdor Palm Oil
Mill near Sungai Bakap plantation (Seberang Prai, Malaysia) in mid
2012. The OPF leaves were removed and the strands were chipped

OH
reducing
sugars

OH

anthraquinone

anthrahydroxyquinone

Fig. 1. Anthraquinone reduction into anthrahydroquinone by reducing sugars.

into small pieces. After sun dried for 3 days, the chips were then
ground to a 1e3 mm size using Wiley mill and the ber was further
dried in an oven at 50  C for 24 h. The OPF biomass was rst subjected to Soxhlet extraction with ethanol/toluene (2:1, v/v) for 6 h
before use. All chemical reagents used in this study were purchased
from Sigma Aldrich, Merck, QRec (Malaysia) and VWR (France) and
used as received. Dried matter contents were determined using a
moisture balance, KERN MRS 120-3 Infra-red moisture analyzer
(drying at 105  C to constant weight). The effective dry matter
content of raw OPF biomass was ~89%.
2.2. Pretreatment with organic scavengers and autohydrolysis
pretreatment
About 20 g (oven dried matter) of OPF were immersed in 100 mL
of acetone containing 0.8 g of 1,8-dihydroxyanthraquinone (Merck)
at room temperature. After thorough mixing, biomass samples
were air-blown to dryness at room temperature followed by
autohydrolysis. The treated OPF sample was loaded into a 0.6 L
stainless steel pressure Parr reactor with a Parr 4842 temperature
controller (Parr Instrument Company, Moline, IL) and was supplemented with an appropriate amount of deionized water to obtain a
nal solid to liquid ratio of 1:9, taking into account the moisture
content of the sample. The mixture was heated at 150  C with
continuous stirring for 8 h (time zero was set when the preset
temperature was reached with the heating rate of 5  C min1,
severity factor: S0 ~4.1). At the end of each reaction, the reactor
was cooled and the liquid phase was recovered by ltration through
Whatman No. 4 lter paper.
2.3. Lignin extraction through organosolv pulping
OPF (200 mm particle size, 25 g dry weight of autohydrolyzed
treated biomass) was mixed with water:ethanol (35:65, v/v) and
0.5% w/w sulfuric acid as a catalyst at 190  C for 60 min (severity
factor: S0 ~2.5), following the method outlined by El Hage et al.
[16]. The solid to liquid ratio used was 1:8. Treatments were carried
out in a 0.6 L stainless steel pressure Parr reactor with a Parr 4842
temperature controller (Parr Instrument Company, Moline, IL). The
reaction mixture was heated at a rate of 5  C min1 with continuous
stirring. At the end of the treatment, the free liquid was removed
and then the brous residue was washed with 89% v/v aqueous
ethanol (3  50 mL) at 60  C and air dried overnight. The washed
liquid samples were combined, and three volumes of water were
added to precipitate the ethanol organosolv lignin (EOL), which
was collected by centrifugation at 4000 rpm for 10 min and then air
dried. The purication of lignin was conducted by extracting it in
the Soxhlet apparatus for 6 h with n-pentane to remove lipophilic
non-lignin matters such as wax, lipids and anthraquinone impurities. The puried organosolv lignin was then dried in an oven at
40  C under atmospheric pressure for another 24 h.
2.4. Characterization of lignin
The FTIR spectrophotometry was carried out in a direct transmittance mode using Perkin Elmer model System 2000 instrument.
The region between 4000 and 400 cm1 with a resolution of 4 cm1
and 20 scans was recorded. The samples were prepared according
to the potassium bromide technique, in a proportion of 1:100
(200 mg of KBr approximately). Interpretation of the IR spectra was
done using Perkin Elmer software.
All nuclear magnetic resonance (NMR) spectroscopy experiments were performed on a Bruker Avance-400 spectrometer.
Quantitative NMR spectra were acquired using an inverse-gated
decoupling (Waltz-16) pulse sequence to avoid Nuclear

M.H. Hussin et al. / Polymer Degradation and Stability 109 (2014) 33e39

Overhauser Effect (NOE). The phosphitilating reagent employed


here was 2-chloro-4,4,5,5-tetramethyl-1,1,3,2-dioxaphospholane
(TMDP) as previously described by Granata and Argyropoulos
[17]. Approximately 25 mg of lignin was added into 400 mL solvent
mixture of pyridine:CDCl3 (1.6/1, v/v) (Aldrich). Then each 150 mL of
chromium (III) acetylacetonate (3.6 mg mL1) and cyclohexanol
(4.0 mg mL1) solution in pyridine/CDCl3 was added to the lignin
solution. The chromium (III) acetylacetonate served as a relaxation
reagent and cyclohexanol as the internal standard. The lignin solution was then vigorously stirred until completely dissolved. Minutes before starting the NMR experiment, approximately 50 mL of
TMDP was added to the vial and the solution transferred into a
5 mm NMR tube. The 31P NMR spectra were recorded with the
following acquisition parameters; inverse-gated pulse sequence,
25 s pulse delay, 200 acquisitions, 61.7 ppm sweep width and 30
pulse width (p1 6.5 usec, pl 1 2.0 db). 13Ce1H 2D Heteronuclear Single Quantum Correlation (HSQC) and Heteronuclear
Multiple Bond Correlation (HMBC) NMR spectra of autohydrolyzed
lignin treated and untreated with 1,8-dihydroxyanthraquinone
were recorded on a Bruker 500 MHz instrument using standard
pulse program.
Lignin samples were subjected to acetylation in order to
enhance their solubility in organic solvents, used in gel permeation
chromatography (GPC). The number average molecular weights
(Mn) and weight (Mw) of lignin were determined after derivatization by GPC with a Dionex Ultimate-3000 HPLC system consisting
of an autosampler and a UV detector and using tetrahydrofuran as
eluent. Standard polystyrene samples were used to construct a
calibration curve. Data were collected and analyzed with Chromeleon software Version 6.8 (Dionex Corp., USA).
The dissolution test was carried out as follows. Approximately
200 mg sample was added to 500 mL distilled water at 25  C and
agitated constantly at 100 rpm for 6 h. 2 mL solution was withdrawn at predetermined intervals and ltered through a 0.45 mm
syringe lter. An equal volume of distilled water was replaced after
each withdrawal [18,19]. After diluted with distilled water, the solution absorbance was measured at 280 nm (Shimadzu UV-2550
Japan). According to the UV linear regression equation, lignin
concentration and the percentage of dissolution (D %) are calculated as follows:

A  A0 K  C  L

(1)

D% Cmax  Vtotal =minitial  100

(2)

where A0 is the absorption of distilled water, A is the absorption of


sample, K is the absorption constant (21 L g1 cm1) [18], C is the
concentration of lignin, Cmax is the maximum concentration at
600 min, L is the thickness of quartz cell (cm), Vtotal is the total
volume and minitial is the initial mass of lignin.
2.5. Measurement of lignin antioxidant activity
2.5.1. Oxygen uptake inhibition
Antioxidant properties of organosolv lignins were investigated
by evaluating oxygen uptake inhibition during oxidation of methyl
linoleate. The induced oxidation by molecular oxygen was performed in a gas-tight borosilicate glass apparatus. Butan-1-ol was
used as solvent for lignin dissolution. Temperature was set to 60  C,
initial conditions inside the vessel were as follows; methyl linoleate
(Fluka, 99%) concentration: 0.32 mol L1; 2,20 -azobisisobutyronitrile (AIBN) (Fluka, 98%) concentration: 7.2  103 mol L1; lignin
concentration: 0.2 g L1; oxygen pressure: 150 Torr. Oxygen uptake
was monitored continuously by a pressure transducer (Viatron
model 104). Without any additive, oxygen uptake is roughly linear

35

and constitutes the control. In the presence of an antioxidant, oxygen consumption is slower, and the antioxidative capacity (OUI) of
organosolv lignin was estimated by comparing oxygen uptake at a
chosen time (3 h), in the presence of this compound (pressure
variation DPsample) and in the absence of the compound (DPcontrol)
according to:

.

DPcontrol  100
OUI% DPcontrol  DPsample

(3)

This ratio denes antioxidative capacity as an oxygen uptake


inhibition index (OUI); it should spread from 0 to 100%, for poor
and strong antioxidants, respectively, and may be negative for
proxidants.
2.5.2. Reducing power assay
The reducing power of samples was determined by the method
proposed by Gulcin et al. [20] after a slight modication. Standard
syringaldehyde and vanillin (Aldrich) solution of concentrations
2.0  102 mg mL1 e 0.1 mg mL1 were prepared. To 1.0 mL of the
standard solution, 2.5 mL of 0.2 M phosphate buffer at pH 6.6
(prepared from the addition of 0.2 M Na2HPO4 and 0.2 M NaH2PO4
(QRec) and 2.5 mL of 1% (w/v) potassium ferricynide, K3Fe(CN)6
(Aldrich) solution were added. The mixture was incubated at 50  C
for 20 min after which 2.5 mL 10% (w/v) trichloroacetic acid
(Merck) was added. The resultant mixture was centrifuged for
20 min at 2500 rpm. The upper layer (2.5 mL) was dispensed and
2.5 mL distilled water and 0.5 mL of 0.1% (w/v) ferric chloride
hexahydrate, FeCl36H2O (Merck) solution were added. The absorbance was measured at 700 nm. The procedure was repeated for
lignin samples.
3. Results and discussion
3.1. Characterization of lignin
Organosolv lignins extracted from oil palm fronds with (DEOL)
and without (EOL) 1,8-dihydroxyanthraquinone as scavenger during the pulping process were characterized. Fig. 2A shows the infra
red spectra of the unmodied and modied organosolv lignin.
Absorption signal at 1160 cm1 which appears in all spectra was
assigned for typical HGS lignin. The presence of absorption bands at
1328 cm1 (syringyl), 1219 cm1 (guaiacyl), 1120 cm1 (syringyl)
and 1033 cm1 (guaiacyl) revealed that OPF lignin were mainly
composed of S and G basic units. In addition the absorbance signal
of S unit (1328 and 1120 cm1) was higher compared to G units
(1272 and 1033 cm1) which suggested the higher syringyl content
[21]. Based on Fig. 2A, when 1,8-dihydroxyanthraquinone (DEOL)
was introduced during the autohydrolysis process, decrease in the
absorption of 1700e1720 cm1 can be observed after the organosolv treatment. These bands may explain the fact that the organic
scavenger reduced signicantly the number of b-keto (Hilbert ketone) groups by condensing with either their precursor hydroxyl
groups or the already formed b-carbonyl groups [22]. Besides, the
reduction was also caused by limitation of the deconstruction on
the lateral chain because of the scavenging effect.
Data from the quantitative 31P NMR of the two organosolv
lignin samples obtained following their derivatization with TMDP
are presented in Fig. 2B. The concentration of each hydroxyl
functional group (in mmol g1) was calculated on the basis of the
hydroxyl content (Table 1) of the internal standard cyclohexanol
and its integrated peak area [17]. From this data, it can be
observed that lignin grafted with 1,8-dihydroxyanthraquinone
(DEOL) gave higher concentrations of HGS phenolic eOH groups
(phenolic eOH 1.49 mmol g1) and lower concentrations of
aliphatic eOH content compared to unmodied organosolv lignin

36

M.H. Hussin et al. / Polymer Degradation and Stability 109 (2014) 33e39

Fig. 2. (A) The infra red and (B) 31P NMR spectra of; unmodied organosolv lignin (EOL) and lignin with 1,8-dihydroxyanthraquinone (DEOL). (H: p-hydroxyphenyl unit; G: guaiacyl
unit; S: syringyl unit).

(EOL) (phenolic eOH 1.35 mmol g1). Very severe conditions


(used during autohydrolysis pretreatment) have enhanced the
cleavage of a- and b-ether linkages between lignin structural units
which later led to the formation of new phenolic eOH when
delignication process takes place [12,13,22]. It was believed that
during the autohydrolysis step, 1,8-dihydroxy anthraquinone was

Table 1
Organosolv lignins characterized by

31

P NMR.

ppm

Assignments ( P NMR)

mmol g1
EOL

DEOL

150e145
144.8
144e140
140e138
138e136
135e133

Aliphatic eOH
Cyclohexanol (internal standard)
Syringyl eOH
Guaiacyl eOH
p-Hydroxyphenyl eOH
Carboxylic acid

1.49

1.08

0.69
0.45
0.21
0.07

0.74
0.46
0.29
0.01

31

reduced in presence of reducing sugars into the highly nucleophilic 1,8-dihydroxyanthrahydroxyquinone (according Fig. 1). The
incorporation of 1,8-dihydroxyanthrahydroxyquinone scavenger
on the lignin matrixes according the well known mechanism [13]
should retard the condensation process (through an electrophilic
substitution reaction at the Ca position) and leads to the production of higher phenolic OH groups. Higher signals for S unit in all
lignin may suggest that OPF lignin contains more syringal basic
unit than guaiacyl unit, as similarly shown in the FTIR spectra.
To acquire further information on the structural characterization of modied organosolv lignin, the autohydrolyzed organosolv
lignin in the presence of 1,8-dihydroxyanthraquinone was subjected to HSQC and HMBC NMR analyses. The main signals in the
aromatic region (dC/dH 100e150/6.0e8.0 ppm) of the HSQC NMR
spectrum are shown in Fig. 3. As previously reported by She et al.
[23], the correlation at dC/dH 103.8/6.69 ppm corresponds to 2/6
position of S units, whereas the correlations for C2eH2 (dC/dH 110.8/
6.96 ppm) and C5eH5 (dC/dH 115.5/6.84 ppm) are assigned to G

M.H. Hussin et al. / Polymer Degradation and Stability 109 (2014) 33e39

37

Fig. 4. 2D-HMBC NMR spectrum at side chain region of lignin with 1,8dihydroxyanthraquinone. (Inlet: Structure of 1,8-dihydroxyanthraquinone).

Fig. 3. 2D-HSQC NMR spectrum of (A) unmodied organosolv lignin (EOL); (B) lignin
with 1,8-dihydroxyanthraquinone (DEOL).

units. Signals for the p-hydroxyphenyl units (H) were identied at


dC/dH 129.7/7.02 ppm. This supports our earlier prediction that the
organosolv lignin obtained from different pretreatments could be
veried as typical HGS lignin. p-coumarates (pCE 2/6) correlations
were also observed in all organosolv lignin samples appearing at dC/
dH 131.2/7.6 ppm. Besides, the presence of anthaquinone ring can be
detected at around dC/dH 120e140/7.0e8.0 ppm respectively.
The modication of the lignin structure in the presence of 1,8dihydroxyanthraquinone also can be supported by the HMBC
analysis (Fig. 4). In this study, a strong interaction at dC/dH 162.0/
4.5 ppm was detected that indicated a strong correlation between
Ca of lignin matrix to C3 and/or C6 position of 1,8dihydroxyanthraquinone. This correlation demonstrates the presence of the anthraquinone structures on the external ring. In general, the acidic conditions during the autohydrolysis and
organosolv treatment lead to the formation of carbonium ion by the
proton-induced elimination of water from benzylic position. The
condensation reactions may occur in the presence of electron-rich
carbon atom such as the C2/C6 presence in guaiacyl and syringyl
ring leading to repolymerization of the lignin. By introducing
organic scavengers like 1,8-dihydroxyanthraquinone, it would bond
to a lignin quinine methide thus forming a covalently-bonded
adduct. The role of anthraquinone here is therefore to induce/
scavenge the carbonium ion ligninelignin condensation reaction
through an electrophilic substitution reaction at the Ca position
[11e13]. Therefore, it is proposed that the modication of lignin in

the presence of 1,8-dihydroxyanthraquinone will form an interaction possibly at C3 and/or C6 of the anthraquinone moiety (Fig. 4).
The weight average (Mw) and number average (Mn) molecular
weight of all lignin were computed from their chromatograms. It
was revealed that the weight average (Mw) of lignin incorporated
with 1,8-dihydroxyanthraquinone ( DEOL: 3624 g mol1) was lower
than the Mw of unmodied organosolv lignin (EOL: 5151 g mol1).
This indicates that the incorporation of organic scavenger (such as
1,8-dihydroxyanthraquinone) with assessment of autohydrolysis
pretreatment facilitates the production of smaller lignin fragments
with low molecular weight. Severe conditions autohydrolysis pretreatment led to more extensive depolymerization (ether linkages
cleavage) of the lignin and decrease the molecular weight. According to Alriols et al. [24], low molecular weight of lignin will
increase its solubility in some organic solvents. In addition, very
low polydispersity (PD) values of lignin grafted with 1,8dihydroxyanthraquinone ( DEOL: Mn 1888 g mol1, PD 1.92)
compared to the unmodied organosolv lignin (EOL: Mn
2373 g mol1, PD 2.17) were observed. The low polydispersity of
lignin contributes to the uniformity of overall packing structures of
lignin and may cause higher solubility [7].
In Fig. 5, we can see that the modied lignin and the unmodied
organosolv lignin exhibit different dissolution curves. The results

Fig. 5. Dissolution proles of unmodied organosolv lignin (EOL) and lignin with 1,8dihydroxyanthraquinone (DEOL).

38

M.H. Hussin et al. / Polymer Degradation and Stability 109 (2014) 33e39

showed that there was about 8.86% of DEOL lignin dissolved,


compared to only 1.15% of unmodied organosolv lignin (EOL). The
solubility of DEOL lignin was 10.67 times higher than the solubility
of EOL. Better solubility of lignin can be obtained when 1,8dihydroxyanthraquinone was incorporated onto lignin structure.
This is due to the fact that 1,8-dihydroxyanthraquinone contains
more polar hydroxyl functional groups thus increase the hydrophilicity of lignin and thereby promotes dissolution in water.
3.2. Lignin antioxidant activity
3.2.1. Oxygen uptake inhibition
The oxygen uptake prole during autoxidation of methyl linoleate is shown in Fig. 6A. By comparing the linear curve of methyl
linoleate alone (control), it appears that all lignin samples exhibited
antioxidant activity by slowing down the oxidation of methyl
linoleate. The antioxidant efciency of lignin has been related with
their structure, phenolic-aliphatic eOH content, purity and polydispersity [25,26]. The assessment of autohydrolysis pretreatment
during the incorporation of organic scavengers has greatly
increased the antioxidant activity of lignin. By referring to the 31P
NMR analysis, it was revealed that the phenolic eOH (syringyl
eOH guaiacyl eOH p-hydroxyphenyl eOH) of lignin grafted
with 1,8-dihydroxyanthraquinone was greater than its aliphatic
eOH. Very severe conditions (used during autohydrolysis pretreatment) have enhanced the cleavage of a- and b-ether linkages
between lignin structural units which later led to the formation of
new phenolic eOH when delignication process takes place [26].
Thus the process generates more phenolic eOH than aliphatic eOH.

In addition, ortho-methoxyl substitution (which was exhibited


in S and G units) might provide resonance stabilization to the
incipient phenoxyl radical that can increase the antioxidant activity
[27,28]. Therefore the antioxidant activity of lignin should increase
with the free phenolic eOH and ortho-methoxyl S and G group
content.
Close
inhibition
values
of
DEOL
(phenolic
eOH 1.49 mmol g1) lignin was observed with OUI ~78% while
unmodied organosolv lignin (phenolic eOH 1.35 mmol g1)
with OUI ~53%. It was also observed that modied lignin that
posses low molecular weight tends to experience high antioxidant
activity than unmodied organosolv lignin (EOL). This observation
is consistent with the previous report regarding the effect of molecular weight of lignin on antioxidant activity [6,26]. The low
molecular weight lignin resulted from the inter-unit bonds cleavage in lignin structure is also accompanied by an increment of
phenolic eOH and decrease of aliphatic eOH.
3.2.2. Reducing power assay
Fig. 6B shows the reducing power of modied and unmodied
organosolv lignin as a function of their concentration. In this assay,
the yellow color of the test solution changes to various shades of
green and blue, depending on the reducing power of each lignin.
The presence of reducers (i.e. antioxidants) causes the reduction of
the Fe3 (ferricyanide) complex to the ferrous complex that has an
absorption maximum at 700 nm [29]. As a comparison, reducing
power of syringaldehyde and vanillin standard was employed to
mimic the reduction ability of syringyl (S) and guaiacyl (G) unit.
Increased absorbance of the reaction mixture indicates increase in
reducing power. Higher reducing power of syringaldehyde than
vanillin standard revealed that the S unit could facilitate better
reduction ability. In this study, the reducing power of lignin
increased with concentration. The reducing power of lignin incorporated with 1,8-dihydroxyanthraquinone (DEOL) was better than
the unmodied organosolv lignin (EOL). It was reported that the
reducing power of some chemical compounds might be due to their
hydrogen-donating ability [30]. Hence higher reducing power of
DEOL was perhaps due to the presence of more electronegative
organic scavenger of 1,8-dihydroxyantharaquinone on the lignin
moiety. Therefore DEOL lignin might contain higher amounts of
reductant, which could react with free radicals to stabilize and
block radical chain reactions. Better antioxidant properties
exhibited by the lignin grafted with 1,8-dihydroxyanthraquinone is
indeed benecial for the various aqueous applications such as mild
steel corrosion inhibition.
3.3. Improved lignin properties for green applications

Fig. 6. Antioxidant prole of unmodied organosolv lignin (EOL) and lignin with 1,8dihydroxyanthraquinone (DEOL); (A) oxygen uptake inhibition and (B) reducing power
method.

Structural alteration of lignin via incorporation of organic


scavenger (1,8-dihydroxyanthraquinone) during the pretreatment
process has tremendously altered both physical and chemical
properties
of
the
lignin.
Lignin
grafted
with
1,8dihydroxyanthraquinone gave higher S-unit with almost no
change to its G-unit. In fact, the presences of G-type unit in the
DEOL lignin structure (with a free C5 position), could give highly
potential active sites for polymerization process such as in phenolformaldehyde condensation reactions. While higher S-type unit (C3
and C5 positions are linked to methoxy group) tends to show a low
reactivity towards formaldehyde [31], but still it offers as a good
antioxidant due to its ortho-methoxyl substitution which might
provide resonance stabilization to the incipient phenoxyl radical
that can increase the antioxidant activity. Higher phenolic eOH and
ortho-methoxyl groups content with smaller fragments of lignin
structures (low average molecular weight) was produced in which
later gave a good antioxidant activity as observed from oxygen
uptake inhibition and reducing power assay.

M.H. Hussin et al. / Polymer Degradation and Stability 109 (2014) 33e39

Furthermore, based from its Mw and PD values, DEOL lignin


could also be used to improve the properties of viscous media used
for inks and paints, extender or phenoleformaldehyde resins
component and for favoring blend capability in polymer formulations. An improvement toward its solubility right after being
modied with 1,8-dihydroxyanthraquinone was observed. The
hydrophobicity of lignin was then reduced with this technique and
therefore this could widen the area of research for lignin to be
applied in aqueous systems. Our current research is concerning the
utilization of modied lignin to be employed as green corrosion
inhibitor.
4. Conclusion
In this work, it was successfully proven that the modication of
the lignin structures by the incorporation of organic scavengers
(1,8-dihydroxyanthraquinone) during the autohydrolysis pretreatment has greatly improved the properties of lignin. Additionally, lignin fractions with higher phenolic hydroxyl content but
lower molecular weight, polydispersity as well as aliphatic hydroxyl content gave higher values of antioxidant activity. Better
solubility
of
lignin
can
be
obtained
when
1,8dihydroxyanthraquinone was incorporated onto lignin structure
due to the fact that 1,8-dihydroxyanthraquinone contains more
polar hydroxyl functional groups thus increase the hydrophilicity of
lignin and thereby promotes dissolution in water. An improved
property of lignin is indeed benecial for its subsequent
applications.
Acknowledgments
The authors are grateful for the nancial support of this research
^le de Compe
titivite
 Fifrom CPER 2007e2013 Structuration du Po
bres Grand'Est and from Universiti Sains Malaysia through USM
Research University Grant e 1001/PKIMIA/854002. M. Hazwan
Hussin would like to express his gratitude to the Ministry of High
Education (MOHE) of Malaysia and Ministere Affaires Etrangeres de
France for the MyPhD scholarship and Boursier du Governement
Francais (CampusFrance). The EA 4370 LERMAB is supported by the
French National Research Agency through the Laboratory of
Excellence ARBRE (ANR-12-LABXARBRE-01).
References
[1] Ragauskas A, Williams C, Davison B, Britovsek G, Cairney J, Eckert C, et al. The
path forward for biofuels and biomaterials. Science 2006;311:484e9.
[2] Sarkar N, Ghosh SK, Bannerjee S, Aikat K. Bioethanol production from agricultural wastes: an overview. Renew Energy 2012;37(1):19e27.
[3] Zhao X, Cheng K, Liu D. Organosolv pre-treatment of lignocellulosic biomass
for enzymatic hydrolysis. Appl Microbiol Biotechnol 2009;82:815e27.
[4] Brosse N, Mohamad Ibrahim MN, Abdul Rahim A. Biomass to bioethanol:
initiatives of the future for lignin. ISRN Mater Sci 2011:1e10. 461482.
[5] Urgatondo V, Mitjans M, Vinardell MP. Applicability of lignin from different
sources as antioxidants based on the protective effects on lipid peroxidation
induced by oxygen radicals. Ind Crop Prod 2009;30(2):184e7.
[6] El Hage R, Perrin D, Brosse N. Effect of pre-treatment severity on the antioxidant properties of ethanol organosolv Miscanthus  giganteus lignin. Nat
Resour 2012;3:29e34.

39

[7] Hussin MH, Rahim AA, Mohamad Ibrahim MN, Yemloul M, Perrin D, Brosse N.
Investigation on the structure and antioxidant properties of modied lignin
obtained by different combinative processes of oil palm fronds (OPF) biomass.
Ind Crop Prod 2014;52:544e51.
[8] Garcia A, Toledano A, Andres MA, Labidi J. Study of the antioxidant capacity of
Miscanthus sinensis lignins. Process Biochem 2010;45(6):935e40.
[9] Vargin MY, Trashin SA, Karyakin A. Corrosion protection of steel by electropolymerized lignins. Electrochem Commun 2006;8:60e4.
[10] Ren Y, Luo Y, Zhang K, Zhu G, Tan X. Lignin terpolymer for corrosion inhibition
of mild steel in 10 % hydrochloric acid medium. Corros Sci 2008;50:3147e53.
[11] Lora JH, Wayman M. Delignication of hardwoods by autohydrolysis and
extraction. Tappi 1979;61:47e50.
[12] Li J, Gellerstedt G. Improved lignin properties and reactivity by modications
in the autohydrolysis process of aspen wood. Ind Crop Prod 2008;27:175e81.
[13] Timilsena YP, Audu IG, Rakshit SK, Brosse N. Impact of the lignin structure of
three lignocellulosic feedstocks on their organosolv delignication. Effect of
carbonium ion scavangers. Biomass Bioenerg 2013;52:151e8.
[14] Blain TJ. Anthraquinone pulping: fteen years later. Tappi J 1993;76(3):
137e46.
[15] Fleming BI, Kubes GJ, Macleod JM, Bolker HI. Soda pulping with anthraquinone. Tappi J 1978;61(6):43.
[16] El Hage R, Brosse N, Chrusciel L, Sanchez C, Sannigrahi P, Ragauskas A.
Characterization of milled wood lignin and ethanol organosolv lignin from
miscanthus. Polym Degrad Stabil 2009;94:1632e8.
[17] Granata A, Argyropoulos DS. 2-chloro-4,4,5,5-tetramethyl-1,3,2-dioxaphospholane, a reagent for the accurate determination of the uncondensed and
condensed phenolic moieties in lignins. J Agric Food Chem 1995;43:1538e44.
[18] Lu Q, Zhu M, Zu Y, Liu W, Yang L, Zhang Y, et al. Comparative antioxidant
activity of nanoscale lignin prepared by a supercritical antisolvent (SAS)
process with non-nanoscale lignin. Food Chem 2012;135:63e7.
[19] Francesco L, Chiara S, Guido E, Francesca M, Giaime M, Anna Maria F. Diclofenac nanosuspensions: inuence of preparation procedure and crystal form
on drug dissolution behavior. Int J Pharm 2009;373:124e32.
[20] Gulcin I, Oktay M, Kirecci E, Kufrevioglu OI. Screening of antioxidant and
antimicrobial activities of anise (Pimpinella anisum L.) seed extracts. Food
Chem 2003;83:371e82.
[21] Hussin MH, Rahim AA, Mohamad Ibrahim MN, Brosse N. Physicochemical
characterization of alkaline and ethanol organosolv lignins from oil palm
(Elaeis guineensis) fronds as phenol substitutes for green material applications. Ind Crop Prod 2013;49:23e32.
[22] Wayman M, Lora JH. Simulated autohydrolysis of Aspen milled wood lignin in
the presence of aromatic additives: structural modications. J Appl Polym Sci
1980;25:2187e94.
[23] She D, Nie XN, Xu F, Geng ZC, Jia HT, Jones GL, et al. Physico-chemical characterization of different alcohol-soluble lignins from rice straw. Cell Chem
Technol 2012;46(3e4):207e19.
[24] Alriols MG, Tejado A, Blanco M, Mondragon I, Labidi J. Agricultural palm oil
tree residues as raw material for cellulose, lignin and hemicelluloses production by ethylene glycol pulping process. Chem Eng J 2009;148:106e14.
[25] Pouteau C, Dole P, Cathala B, Averous L, Boquillon N. Antioxidant properties of
lignin in polypropylene. Polym Degrad Stabil 2003;81(1):9e18.
[26] Pan XJ, Kadla JF, Ehara K, Gilkes N, Sadler JN. Organosolv ethanol lignin from
hybrid poplar as a radical scavenger: relationship between lignin structure
extraction conditions and antioxidant activity. J Agric Food Chem
2006;54(16):5806e13.
[27] Barclay LRC, Xi F, Norris JQ. Antioxidant properties of phenolic lignin model
compounds. J Wood Chem Technol 1997;17(1e2):73e90.
[28] Nadji H, Diouf PN, Benaboura A, Bedard Y, Riedl B, Stevanovic T. Comparative
study of lignin isolated from Alfa grass (Stipa tenacissima L.). Bioresour
Technol 2009;100:3585e92.
[29] Ferreira ICFR, Baptista P, Vilas-Boas M, Barros L. Free-radical scavenging capacity and reducing power of wild edible mushrooms from northeast
Portugal: individual cap and stipe activity. Food Chem 2007;100:1511e6.
[30] Shimada K, Fujikawa K, Yahara K, Nakamura T. Antioxidative properties of
xanthan on the autoxidation of soybean oil in cyclodextrin emulsion. J Agric
Food Chem 1992;40:945e8.
[31] Tejado A, Pena C, Labidi J, Echeverria JM, Mondragon II. Physicochemical
characterization of lignins from different sources for use in phenoleformaldehyde resin synthesis. Bioresour Technol 2007;98:1655e63.

Vous aimerez peut-être aussi