Sei sulla pagina 1di 10

Industrial Crops and Products 84 (2016) 284–293

Contents lists available at ScienceDirect

Industrial Crops and Products


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

Dissolution of kraft lignin using Protic Ionic Liquids and


characterization
Tazien Rashid a , Chong Fai Kait b , Iyyasamy Regupathi c , Thanabalan Murugesan a,∗
a
Department of Chemical Engineering, Universiti Teknologi Petronas, Bandar Seri Iskandar, Tronoh32610, Perak, Malaysia
b
Fundamental and Applied Sciences Department, Universiti Teknologi Petronas, Bandar Seri Iskandar, Tronoh 32610, Perak, Malaysia
c
Department of Chemical Engineering, National Institute of Technology, Surathkal P. O. Srinivasnagar, Mangalore 575 025, Karnataka, India

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

Article history: In the present research three Protic Ionic Liquids (pyridinium formate, pyridinium acetate and pyridinium
Received 24 August 2015 propionate) were synthesized and tested for the dissolution and subsequent regeneration of kraft lignin.
Received in revised form 5 February 2016 Among the investigated solvents, pyridinium formate showed a higher dissolution capacity (70% w/w)
Accepted 6 February 2016
i.e. (710 g/L) at 75 ◦ C within 1 h. The results indicated that the introduced solvent is thermally stable,
noncorrosive, possesses low viscosity and is easy to recycle. The dissolution process is purely physical
Keywords:
and the physicochemical analysis of the regenerated lignin showed high thermal stability, with reduction
Lignocellulose
in polydispersity and the average molecular weight was reduced from 4119 g/mol to 1249 g/mol. FTIR
Kraft lignin
Protic Ionic Liquids
spectroscopy and 1 H NMR results proved that the regenerated lignin is less degraded. Moreover the O H
Dissolution vibrations of regenerated lignin showed a weak inter and intramolecular interaction in regenerated lignin,
Regeneration which could positively help in reducing its chemical resistance towards processing for further commercial
applications. Due to the higher solubility of lignin and its stability towards recyclability, the pyridinium
formate proved that present selective dissolution and regeneration of lignin could significantly enhance
the pretreatment techniques for lignocellulosic biomass.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction cient, cost effective technology and selective separation of native


lignin. Kraft and sulfite processes are the oldest and most common
Lignin is considered as the second most abundant organic, nat- technologies for commercial delignification of wood. However, the
ural compound in the world and is also considered as a renewable high energy inputs and potential pollutants involved (sulfur con-
source of aromatic and chemical feedstock. The concept of bio refin- taining reagents) in the process make it uneconomical. Moreover
ery comprises the selective separation of the three main polymeric high temperature and pressure conditions degrade the cellulose
wood components namely, cellulose, hemicellulose and lignin, and structure and lead to the destruction of the fermentable sugar i.e.,
their subsequent utilization for the production of fuels and high glucose (Baptista et al. 2008; Gellerstedt and Henriksson, 2008; Vila
value added chemicals. Cellulose and hemicellulose fractions can be et al., 2003). Organosolv pulping process has several advantages
readily used as a starting material for the production of biofuels and over sulphite and kraft process through which some of the above-
biochemicals, whereas the efficient utilization of lignin presents mentioned drawbacks were eliminated (Alriols et al., 2009; Li et al.,
an ongoing challenge (Zhang, 2008). The separation techniques 2012; Vila et al., 2003). Though it is possible to obtain a sulfur free
employed for lignocellulosic biomass components (i.e., cellulose, lignin with minimum degradation, however the major drawbacks
hemicellulose, and lignin) are mostly conducted in destructive of this technique are thermal instability of solvent and high cost
ways to obtain comparatively pure cellulose, which results in for solvent regeneration (Sundquist, 1988). da Costa Lopes et al.
degraded lignin that can only be used as a low value by products or (2013a) introduced a fractionation methodology by using 1-ethyl-
burnt as a low-grade fuel (Zhang, 2008; Zakrzewska et al., 2010). 3-methylimidazolium acetate for wheat straw cooking followed by
Therefore a key challenge remains on the achievement of an effi- alkaline extraction at 120 ◦ C for 6 h and the reported lignin contents
were very low (i.e., 41.8% cellulose, 25.4% hemicellulose and 8.0%
lignin). Later a multiple linear regression model was developed by
da Costa Lopes et al. (2013b) based on their experimental results at
∗ Corresponding author. Fax: +60 53656176.
various operating conditions.
E-mail addresses: murugesan@petronas.com.my, tmgesan 57@yahoo.com
(T. Murugesan).

http://dx.doi.org/10.1016/j.indcrop.2016.02.017
0926-6690/© 2016 Elsevier B.V. All rights reserved.
T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293 285

Lignin is the second most important constituent of lignocel-


lulosic biomass, although far less attention has been given to its
recovery. The lignin content in normal hardwood and softwood
varies from 20% to 30% and 26% to 32% respectively (Maki Arvela
et al., 2010; Sun and Cheng, 2002). Lignin is a highly branched aro-
matic polymer that binds cellulose and hemicellulose via strong
hydrogen bonding and ester linkages, which act as “glue binding”
in the whole lignocellulosic biomass. These linkages and molecular
interactions cause rigidity and microbial resistance in the ligno-
cellulosic biomass. Furthermore, lignin behaves quite differently
Fig. 1. General scheme of the reaction.
in solutions compared to cellulose, as the dissolution mechanisms
for lignin and cellulose are different. The solubility of cellulose
increases almost linearly with hydrogen bonding strength, but in
case of lignin it is reverse (Hart et al., 2015; Horvath, 2006; Lee et al.,
2009; Vitz et al., 2009).
Recently Immidazolium based Aprotic Ionic Liquids (AIL’s) with
short side alkyl chains have been extensively used for the dis-
solution and delignification of biomass (King et al., 2009; Pu
et al., 2007). A comprehensive studies regarding lignocellulose
(i.e., cellulose, lignin, and monosaccharides) solubility in IL’s was
performed by Zakrzewska et al. (2010). Among the IL’s investi-
gated, Immidazolium chloride was found to be more suitable for
cellulose dissolution, whereas Immidazolium acetate was consid-
ered as the best IL for lignin dissolution. It was also concluded
that acetate as anion reduces the thermal stability of the sol-
vent. Despite their usefulness, AIL’s do have certain drawbacks
namely high viscosity and high operating temperature and recov-
erability etc. Apart from these issues, extended dissolution times
are required for processing, (generally >12 h). High viscosity and
1
cost of IL’s were reduced by using polar organic solvents as a Fig. 2. H NMR spectra of pyridinium formate and its precursors in DMSO-d6.

co-solvent and a variety of co-solvents, including dimethylsul-


foxide (DMSO) and N,N-dimethylformamide (DMF), have been mineral acids commonly used for the delignification of the ligno-
tested for delignification (Mai et al., 2014; Pinkert et al., 2011; cellulose (Li et al., 2012; Zhang et al., 2010; Zhou et al., 2012).
Rinaldi, 2011). However a comprehensive study on the effect of
co-solvents is still lacking. Protic Ionic liquids (PIL’s) possess excel-
lent chemical and thermal stability and negligible vapor pressure 2. Material and methods
and they are capable of hydrogen bonding, including proton accep-
tance and proton donation. Protic IL’s are less expensive than 2.1. Materials
traditional Immidazolium-based AIL’s. Protic acetate PIL’s were
studied for delignification of biomass and the lignin extraction was The chemicals used for the present research namely, Pyri-
found to be ≥50% w/w. But due to a lower degree of protonation dine, Formic acid, Glacial Acetic acid, Propionic acid, Chemicals for
(e.g., less ionicity) of amines by acetic acid protic acetates were Karl Fischer titration, Dimethyl Sulphoxide (DMSO-d6 ) (used as a
found to be thermally unstable (Achinivu et al., 2014). Recently solvent for NMR samples), Lignin (kraft lignin-indulin AT), Micro-
a techno-economic investigation study along with thermal sta- crystalline Cellulose (MCC) (with a particle size of 20–150 ␮m),
bility comparison was reported for protic ammonium hydrogen Sodium hydroxide, Methanol, Acetone (99% purity), were of ana-
sulfate using AIL (i.e., 1-ethyl-3-methylimidazolium acetate) as lytical grade and purchased from Sigma–Aldrich. All the chemicals
benchmark and the results are comparable (George et al., 2015). were used as received. Triple distilled water was used for the prepa-
Due to their numerous desirable properties they offer an attractive ration of all aqueous solutions.
alternative to conventional nonionic solvents. Based on the disad-
vantages of the available techniques for complete delignification 2.2. Synthesis of pyridinium based Protic Ionic Liquids
and regeneration of lignin, in the present research an attempt has
been made to use cost effective and easy to synthesize pyridinium Pyridinium based Protic Ionic Liquids are produced when a pro-
based Protic Ionic Liquids with different anionic compounds and ton is transferred from a carboxylic acid to pyridine and the general
hydrogen bond basicity for the selective dissolution of lignin. scheme of the reaction is shown in Fig. 1. For the present study
For the present research pyridine is selected due to its classic three anions are selected based on increasing alkyl chain length
tailored property, through which pyridine can be easily attached i.e. [HCOO− ], [CH3 COO− ] and [CH3 CH2 COO− ], and the PIL’s namely
to different anions (Scriven and Murugan, 2000) also it has a high pyridinium formate [PyFor], pyridinium acetate [PyAce] and pyri-
mobile proton with a tendency of both electron pair donor and dinium propionate [PyPro] respectively were prepared according
proton acceptor (Shimizu et al., 2000). Pyridine is used in drug to the established procedures (Belieres and Angell, 2007). The PIL’s
formulation, insecticides, plant growth regulator, and other agri- thus prepared were kept in the vacuum oven at 80 ◦ C for 48 h, in
cultural products etc. (Chaubey and Pandeya, 2011; Haviv et al., order to remove the excess moisture, which was formed during
1983; Maga, 1981). In this study carboxylic acid anions are prefer- the reaction. The oven dried solvent was sealed with laboratory
ably selected due to their advantages; namely, low melting points, parafilm to prevent any moisture contamination. The characteri-
low viscosities, and high hydrogen bonding acceptor abilities, all of zation analysis using 1 H NMR (DMSO-d6 , 500 MHz) (Fig. 2.), TGA
which should facilitate the dissolution of lignin. Moreover, weak (Fig. 3.) and Karl Fischer water content analysis performed on the
organic acids are less toxic and less corrosive compared to the [PyFor] showed the following results respectively; [ı = 7.37(2H, d),
286 T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293

of 1:200, in an agate mortar and the resulting mixture was pressed


at 7 tons for 30 s to form pellets. These pellets were analyzed using
32 scans at a resolution of 2 cm−1 . Four scans were recorded for
each FTIR spectra and blank KBr pellets were used as the standard
background.

2.5.2. 1 H Nuclear magnetic resonance spectroscopy (NMR)


The one-dimensional (1D) NMR spectra was measured using a
Bruker AVANCE 400 NMR spectrometer at an operating frequency
of 500 MHz and room temperature. The samples were dissolved
in DMSO-d6 . The 1 H chemical shifts were determined in parts per
million.

2.5.3. Gel permeation chromatography (GPC)


Agilent 1200 Series gel permeation chromatography (GPC) sys-
tem (dual UV/vis detector with wavelength, 235 nm and 254 nm)
was used for all experiments. Agilent 1200 Series refractive index
detector with an automatic recycle valve was used to analyze the
samples. A calibration curve was created based on 12 polystyrene
standards (PSS Ready Cal-Kit Poly) from molecular weight of
Fig. 3. TGA thermograms of [PyFor] and its precursors.
266 g/mol to 2.52 × 106 g/mol. Data collection and processing were
done by Agilent ChemStation with GPC-SEC data analysis software.
7.77(1H, t), 8.16 (1H, s), 8.57 (2H, t), 9.75(1H, s); Thermal decom-
position temperature Td = 115 ◦ C; Water contents = 451 ppm]. 2.5.4. Thermogravimetric analysis (TGA)
A PerkinElmer, Pyris V-3.81 thermal gravimetric analyzer was
2.3. Solubility test for kraft lignin used to measure the onset and thermal decomposition tempera-
ture “Td ” of the samples. The equipment was heated at a rate of
Lignin solubility tests were conducted by dissolving 10% (w/w) 10 ◦ C min−1 with a temperature range of 50–800 ◦ C. Approximately
of the lignin in glass vials containing pyridinium solvent with dif- 4 mg of the samples were placed in an aluminium pan and the
ferent anions i.e., formate, acetate and propionate to determine samples were analyzed under N2 gas blanket using a flow rate of
ideal combination for highest lignin solubility. Choice of anions was 20–25 ml/min.
based on increasing alkyl chain length and hydrogen bond basicity
(Hart et al., 2015). The vials were sealed with parafilm M (labo- 2.5.5. Differential scanning calorimetry (DSC)
ratory film) and the mixture was then placed into silicon oil bath A Waters DSC Q2000 V24.11 Build 124A instrument calibrated
and heated with vigorous magnetic stirring (700 rpm). The desired with Indium was used to measure the glass transition tempera-
temperatures for the dissolution experiments were monitored and ture “Tg ” of the samples. All the samples were dried at 150 ◦ C at
measured (±1 ◦ C). The solutions were then visually checked for the a rate of 20 ◦ C min−1 and then cooled and equilibrated at 20 ◦ C
complete solubility. If a given solution was transparent with no before measurements. Approximately 7 grams of samples were
undissolved solids, the amount of the lignin was then incrementally carefully placed in an aluminium crucible. The heating rate during
added to increase the mass fraction until the saturation point was the measurement was 5 ◦ C min−1 . All reported data are the average
achieved. Apart from visual observations the saturation point was of duplicates.
also observed and confirmed using Meiji optical microscope MT
4000 at 100× and 4× (Glas et al., 2015). The concentrations of lignin 3. Results and discussions
in the solution were measured by UV–vis spectrophotometer using
the stock solutions of known concentrations (150–1350 ppm). The 3.1. Effect of anion and alkyl chain length on solubility of kraft
concentration of unknown samples were determined by using lignin
the standardized calibration curves made for each of the solvents
namely [PyFor], [PyAce] and [PyPro] (Wang et al., 2011). Dissolution of commercial kraft lignin in pyridinium based sol-
vents with different anions and hydrogen bond basicity is studied.
2.4. Solubility test for cellulose The investigated anions were formate, acetate and propionate. The
solubility of lignin in pyridine and the respective precursors at dif-
Cellulose solubility tests were conducted by preparing 15% ferent temperatures is shown in Fig. 4. It was observed that the
(w/w) sample of cellulose in [PyFor] and its precursors. The vials lignin solubility was maximum i.e., 26% (w/w) in case of formic
were placed on silicon oil bath and stirred for different time inter- acid (at ≈ 50 ◦ C) after that it became approximately constant. Acetic
vals. The desired temperatures were monitored and measured acid showed a very little solubility i.e., ≤1% (w/w) and for propi-
(±1 ◦ C). The sample was vacuum filtered and then the weight of onic acid it was almost negligible at all temperatures. Based on
cellulose dissolved was determined gravimetrically. The similar these observations it can be concluded that the lignin solubil-
procedure was extended for both pyridinium acetate [PyAce] and ity decreases with increasing alkyl chain length of the precursors
pyridinium propionate [PyPro] solvents. for pyridinium based compounds, whereas the maximum lignin
solubility in pyridine was 47% (w/w). The calculated solubility
2.5. Characterizations using UV–vis spectrophotometer is reproducible within a devia-
tion of <1% and closely matches with the experimental solubility
2.5.1. Fourier transform infrared analysis (FTIR) determined by microscopic observations. The solubility of lignin
The FTIR spectra were taken using PerkinElmer spectrome- in [PyFor], [PyAce] and [PyPro] are compared in Fig. 5. For the
ter, applying the wave number in the domain of 5000 cm−1 to case of the present prepared pyridinium formate [PyFor] the lignin
1000 cm−1 . Recovered lignin was mixed with KBr in a weight ratio solubility was found to be ≥70% (w/w) at 75 ◦ C which is clearly
T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293 287

Table 1
Comparison of reported solubilities of kraft lignin in various solvents and present research.

Solvent type Solvent name Raw material Conditions Lignin solubility (g/L) References

Organic solvent Ethanol/water system ALCEL lignin 23 ◦ C, 24 h 166 Ni and Hu (1995)


Ionic liquid 1,3-dimethylimidazolium methylsulfate, 75 ◦ C,24 h 344 Pu et al. (2007)
Ionic liquid 1-hexyl-3-methyimidazolium Kraft lignin 75 ◦ C,24 h 275 Pu et al. (2007)
trifluoromethanesulphonate
Ionic liquid 1-propyl-3-methylimidazolium bromide Alkali, low 80–90 ◦ C, 20 min 62.4 Lateef et al. (2009)
sulfonate content
lignin
Ionic liquid 1-ethyl-3-methylimidazolium acetate Kraft lignin 90 ◦ C, 24 h >300 Lee et al. (2009)
Organic solvent 1,4-Butanediole/water system Kraft lignin 25 ◦ C, 4 h 14 Wang et al. (2011)
Ionic liquid Ammonium, phosphonium and pyrrolidinium Kraft lignin 90 ◦ C,3 h 460–390 Glas et al. (2015)
based ionic liquids
Protic Ionic Liquid Pyridinium formate Kraft lignin 75 ◦ C, 1 h 710 Present work

to acetate [Ace] and propionate [Pro] anions with comparatively


higher pka values i.e., 4.76 and 4.88 respectively. This could also
indicate that lignin is more soluble in the solvent with a higher
degree of protonation (e.g., more ionicity). The low hydrogen bond-
ing basicity of anion and the shorter alkyl chain length is favorable
for lignin dissolution as reported by (Hart et al., 2015; Glas et al.,
2015). The lignocellulosic components i.e., cellulose, hemicellulose
and lignin are interlinked through hydrogen and covalent bond-
ing (Remsing et al., 2008). The anionic part of the solvent plays
an important role during dissolution process which breaks the
intramolecular hydrogen bonding network in lignocellulose, thus
allowing the dissolution of lignin. It can be easily realized that
cationic part is same for the three reported solvents while the dif-
ference is only in the anionic part. In this case the anionic part
with low hydrogen bond basicity and smaller alkyl chain length
would be a choice for selective dissolution of lignin as shorter the
alkyl chain length, higher is the lignin solubility. It is evident that
the high hydrogen bonding strength that has been reported to be
important for cellulose dissolution is not as influential for the case
of lignin (Muhammad et al., 2013). Based on the recent literature
Fig. 4. Solubility of kraft lignin in precursors. (Glas et al., 2015; Lateef et al., 2009; Lee et al., 2009; Pu et al.,
2007), the reported solubilities of kraft lignin in different ionic liq-
uids were less than 50% (w/w) in all the cases whereas the present
solvent, pyridinium formate [PyFor] showed a very high percent-
age of lignin solubility (70% w/w) at 75 ◦ C within 1 h (Table 1). The
reported solubilities in IL’s (Table 1) were measured at relatively
higher temperatures (≈90 ◦ C). Based on the above results, since the
observed lignin solubility is higher in [PyFor] solvent (70% w/w),
further experiments were conducted using [PyFor] only.

3.2. Effect of water contents on kraft lignin solubility

The preparation of pyridinium based solvents involves an acid


base neutralization reaction where water is produced as one of the
product and hence the effect of water contents is an important
parameter of interest. The effect of water content on the solubil-
ity of lignin was conducted by deliberately adding triple distilled
water (10–95% (v/v)) to the pure [PyFor] solvent. The solubilities
of lignin in aqueous solutions of [PyFor] are shown in Fig. 6. It can
be seen that the solubility of lignin is highly dependent on water
content and showed a decreasing trend with an increase in the
water content. Lignin solubility was very low and almost negli-
Fig. 5. Solubilities of kraft lignin in pyridinium formate [PyFor], pyridinium acetate gible at higher concentrations of water. For the cases where the
[PyAce] and pyridinium propionate [PyPro].
water content is above 50% (v/v) the solubility becomes negligi-
ble. Swatloski et al. (2002) reported that increased water contents
much higher than [PyAce] and [PyPro]. The lignin solubility fol- (i.e. >1% (10,000 ppm)) would impair the dissolution capacity of the
lows the order: [PyFor] (70% (w/w) ≥ [PyAce] (64% w/w) ≥ [PyPro] solvent as water would compete with anionic part of the solvent
(55% w/w). [PyFor] dissolves the maximum amount of lignin among and crowd the hydrogen bond accepting sites of anion, which leads
the other solvents studied in the present research (Fig. 5). Among to difficult interaction of anion to lignin and hence the lignin sol-
the anions tested, the formate anion has the lowest pka value of ubility gets reduced (Swatloski et al., 2002). Few researchers have
3.77 and showed the maximum dissolution for lignin as compared reported that the water contents at the level of parts per million
288 T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293

Fig. 6. Effect of water content in [PyFor] on solubility of kraft lignin.

Fig. 7. (a) Comparison of FTIR spectrum of kraft lignin and PyFor] regenerated lignin.
does not have major impact on lignin solubility (Ni and Hu, 1995; (b) Comparison of O–H vibrations of kraft lignin and [PyFor] regenerated lignin.
Pinkert et al., 2011; Wang et al., 2011).

3.3. Regeneration of kraft lignin matic rings are transformed into quinonoid structures. In addition,
the intensities of the bands corresponding to the phenolic hydroxyl
Regeneration of dissolved kraft lignin from the solution was content are slightly lower in [PyFor] treated lignin i.e., 1382 cm−1
made using water as an antisolvent (Glas et al., 2015). The dissolved compared to kraft lignin. The lower amount of phenolic hydroxyl
lignin in [PyFor] was taken in a 100 ml flask with 40 ml of water. The contents in lignin indicates a less degraded structure as reported
flask was sealed and stirred for complete mixing of the contents. by Gellerstedt and Henriksson (2008) which can be useful for fur-
Lignin which is insoluble in water precipitates out and separated by ther potential lignin applications such as phenolic compounds and
vacuum filtration by applying minimum vacuum. Repeated wash- carbon fibers.
ing of the precipitates was carried out with triple distilled water For the O H vibrations of lignin, an extensive vibration band
and acetone to ensure that all the solvent has been washed out lies in the range associated with 3200–3500 cm−1 as reported by Ji
with water and collected in the beaker. The collected lignin was et al. (2012). The absorption band of the O H vibration of the kraft
then oven dried at 70 ◦ C overnight and the amount of lignin recov- lignin is 3375 cm−1 , while that of the regenerated lignin shifts to a
ered was estimated gravimetrically. On the other hand the collected higher wavenumber 3384 cm−1 (Fig. 7b), which could be attributed
solvent and the wash water mixture were saved for recycling pur- to the reason that during dissolution and regeneration process the
pose. It was observed that lignin recovery was nearly same (≥95%) intramolecular interactions of lignin are weakened. This week inter
for all temperatures (i.e., 25 ◦ C, 50 ◦ C, 75 ◦ C, and 100 ◦ C) studied in and intramolecular interaction in recovered lignin could positively
the present research. Even though a very slight decrease in recov- affect its heterogeneity and can help in reducing its chemical resis-
ery was observed at high temperature conditions i.e., at 125 ◦ C, the tance towards processing of this feedstock for further commercial
recovered lignin however remains chemically unchanged. applications. The results of FTIR spectra of kraft lignin and regen-
erated lignin are comparable and there are no new peaks appeared
in the spectrum, which indicates that there is no chemical reac-
3.4. Analysis of regenerated lignin
tion during the dissolution and regeneration process of lignin and
[PyFor] can safely considered as a physical solvent for lignin.
3.4.1. FTIR analysis
FTIR analysis was performed for the [PyFor] treated and recov-
ered lignin samples using commercially available kraft lignin as 3.4.2. Nuclear magnetic resonance spectroscopy
a standard. The FTIR spectra for the presence of hydroxyl, car- The 1 H NMR analysis was performed for the [PyFor] treated and
bonyl, methoxyl, and carboxyl functional groups respectively were recovered lignin samples using commercially available kraft lignin
analyzed according to literature that have been reported for the as a standard. The 1 H NMR spectra for the presence of aromatic
identification of lignin (Table 2). FTIR analysis of the present [PyFor] protons in guaiacyl and syringyl units respectively were analyzed
treated and regenerated lignin is shown in Table 3. Lignin molecule according to literature for the identification of lignin (Table 4. The
contains various functional groups such as O H, C H, and C O major peaks of aromatic protons in guaiacyl propane (G) appear
etc. These functional groups produce specific transmittance at spe- between 7.5 and 6.7 ppm and similarly that of the aromatic protons
cific wave numbers. FTIR spectra of kraft lignin and [PyFor] treated in syringyl propane (S) appear between 6.7 and 6.1 ppm (Mainka
lignin are compared in Fig. 7a and it was observed that at all spe- et al., 2015; Zhang et al., 2010; Zhou et al., 2012). The shift of
cific intensities reported for [PyFor] treated lignin produced low signals at 2.7 ppm corresponds to DMSO-d6. The chemical shift
intensity of aromatic and guaiacyl (G) rings (Table 3). For [PyFor] of methoxyl protons ( OCH3 ) shows a sharp signal at 3.7 ppm in
treated lignin the observed intensities of the bands corresponding kraft lignin and [PyFor] regenerated lignin. Protons in aliphatic
to the aromatic skeletal vibration and guaiacyl (G) ring breathing and hydrocarbon region of lignin produce peaks between 2.0 and
are 1510 cm−1 and 1266 cm−1 respectively. These values are lower 0.8 ppm. It was observed that peak assignments in case of [PyFor]
than those reported for kraft lignin, which could be attributed to the regenerated lignin are comparable to that of kraft lignin which indi-
reason that during dissolution and regeneration processes the aro- cated that there are no chemical reactions between [PyFor] and
T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293 289

Table 2
Band assignments in kraft lignin, hardwood and softwood lignin for FTIR spectra.

Band (cm−1 ) Assignment

Kraft lignin Softwood lignin Hardwood lignin

3200–3500 O H vibrations
1725 1735 C O stretching (unconjugated)
1660 1658 C O stretching (Conjugated)
1598 1596 1603 Aromatic skeletal vibration breathing with C O stretching
1513 1510 1510 Aromatic skeletal vibration, Ga > Sb
1463 1463 1462 C H deformations asymmetric
1426 1423 1425 Aromatic skeletal vibrations combined with C H in-plane deformation
1384 1375 1375 Phenolic OH and aliphatic C H in methyl groups
1328 S unit breathing with C O stretching and condensed G rings
1269 1269 1269 G ring breathing with carbonyl stretching
1218 1221 1220 C C plus C O plus C O stretch; G condensed > G etherified
1139 1140 1140 C H in-plane deformation of G ring plus secondary alcohols plus C O
stretch
1126 Ether–O–
1116 Aromatic C H deformation in S ring
1086 1086 1086 C O deformation in secondary alcohols and aliphatic esters
1032 1032 1033 Aromatic C H in-plane deformation (G > S) plus C O deformation in
primary alcohols plus C O stretch (unconjugated)
a
Guaiacyl unit.
b
Syringyl unit.

Fig. 8. Comparison of 1 H NMR spectra of kraft lignin and [PyFor] regenerated lignin.

bonds in lignin molecules indicating the dissolution process as pure Mn , (ii) the weight average molecular weight Mw and (iii) poly-
physical process. dispersity index (PDI) are of common interest. Ratio of (Mw /Mn )
is the polydispersity index (PDI), which describes the overall dis-
3.4.3. Molar mass distribution by gel permeation tribution of the molar masses in the lignin (Erdocia et al., 2014;
chromatography (GPC) Pinkert et al., 2011; Santos et al., 2014). The results show that the
Gel permeation chromatography (GPC) is a technique that [PyFor] treated lignin has lower molecular weights than that for the
provides insight into the molecular weight distribution and frag- pure kraft lignin (Table 5). The reduction in molecular weight indi-
mentation of lignin during dissolution and regeneration process. cates that the lignin molecule has been fragmented into smaller
Low polydispersity is preferred which increases the possibility of units, during dissolution process. The molecular weights and the
softening for making carbon fiber composites from lignin (Kubo polydispersity of the lignin recovered from [PyFor] treatment are
and Kadla, 2005). Changes in the average molecular weight of the smaller than that for the kraft lignin samples indicating that the
lignin were determined and compared for the [PyFor] treated lignin regenerated lignin has a more homogenous composition. The PDI
and kraft lignin. Gel Permeation Chromatography (GPC), a common is low for [PyFor] treated lignin which is an added advantage for
method was used to determine molar mass. The lignin samples further processing of lignin. This finding is further supported by
were fully dried under vacuum at 40 ◦ C and dissolved in tetrahy- the SEC chromatographs from which the molecular weights of
drofuran (THF). The molecular weight was then characterized by samples were calculated, Fig. 9 shows that for the [PyFor] treated
GPC using size exclusion chromatography (SEC). Three character- lignin the intensity of the peak at >8 min is heightened. This high
istics of lignin namely (i) the number average molecular weight intensity of [PyFor] treated lignin confirms that the dissolution and
290 T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293

Fig. 9. Size exclusion chromatograph of kraft lignin and [PyFor] regenerated lignin.

Fig. 10. Thermogravimetric analysis of kraft lignin and [PyFor] regenerated lignin.
Table 3
FTIR analysis of [PyFor] regenerated lignin.

Kraft lignin Regenerated lignin from [Py][For]


−1
Band (cm ) Band (cm−1 )
3375 3384
1600 1600
1598 1596
1513 1510
1463 1461
1426 1425
1384 1382
1269 1266
1218 1215
1139 1138
1086 1086
1032 1031

Table 4
1
H NMR assignments of kraft lignin and [PyFor] regenerated lignin.

Range number Peak, ppm Main assignment

1 7.5–6.7 Aromatic protons in guaiacyl units


2 6.7–6.1 Aromatic protons in syringyl units Fig. 11. DTGA thermograms of kraft lignin and [PyFor] regenerated lignin.
3 6.1–5.8 H␣ of ␤-O-4 structures
4 5.8–5.2 H␣ of ␤-5 structures
5 5.2–4.9 H of xylan residue
3.4.4. Thermal decomposition behavior
6 4.9–4.3 H␣ and H␤ of ␤-O-4 structures
7 4.3–4.0 H␣ of ␤-␤ structures and H of xylan residue Thermal Gravimetric Analysis (TGA) is a convenient method
8 4.0–2.5 H of methoxyl protons ( OCH3 ) to measure the weight loss with increasing temperature over a
9 2.5–2.2 H of aromatic acetates definite period of time. The thermal stability of the kraft lignin
10 2.2–1.6 H of aliphatic acetates and the regenerated lignin was examined by Thermal Gravimet-
11 1.6–0.8 Hydrocarbon protons
ric Analysis (TGA) and first derivative thermogravimetric analysis
(DTGA) thermograms. The TGA and DTGA curves of kraft lignin
Table 5
and the regenerated lignin from [PyFor] solution are shown in
Weight average (Mw ) and number-average (Mn ) molecular weights (gmol−1 ) and Figs. 10 and 11 respectively. The results indicated low onset tem-
polydispersity (Mw /Mn ) of the treated and untreated lignin. perature for kraft lignin (curve A) at 353 ◦ C, and the onset was
Solvent Mn (g/mol) Mw (g/mol) PDI (Mw/Mn)
increased to 390 ◦ C for the regenerated lignin. It can be due to
the development of the condensed linkages of regenerated lignin,
Pure kraft lignin 1249 4119 3.29
which is extremely stable. However the overall thermal decompo-
Pyridinium formate—treated lignin 902 1738 1.92
sition behavior of kraft lignin and [PyFor] treated lignin was found
to be similar.
For the DTGA the thermal decomposition behavior of lignin can
regeneration with [PyFor] can produce more uniform lignin frag- be divided into three major steps for the weight loss in the present
mentation with low degree of polydispersity. These results are in studied temperature range of 25–800 ◦ C. Step I (25–150 ◦ C) involves
agreement with the standard methods for the determination of the the dehydration of lignin which ensures the complete removal
molar mass distribution of lignin reported by Baumberger et al. of water and other low molecular weight volatiles, whereas step
(2007).). The 1 H NMR spectra of kraft lignin and [PyFor] treated II (151–565 ◦ C) involves the fragmentation of inter-unit linkages
lignin are compared in Fig. 8 of lignin, the primary products of this step are coke, organic and
T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293 291

Fig. 12. DSC thermograms of kraft lignin and [PyFor] regenerated lignin.
Fig. 14. Effect of lignin contents in the presence of MCC on lignin recovery.

Fig. 13. Solubility of microcrystalline cellulose (MCC) in pyridinium formate [PyFor]


and its precursors.
Fig. 15. Lignin solubility in pure and recycled pyridinium formate [PyFor].

phenolic compounds, and gaseous products. In step III (566–800 ◦ C)


the thermal decomposition of the aromatic rings of lignin takes
place and char or coke are formed in this range (Sun et al., 2013;
kraft lignin respectively. These results are comparable to those
Tejado et al., 2007; Vasile et al., 2006). The maximum weight loss
reported in literature, where the reported Tg was in the range of
rates appeared at step II at 280 ◦ C at a rate of 1.49% min−1 for kraft
80 and 180 ◦ C (Tejado et al., 2007; Vasile et al., 2006). No signif-
lignin and at 340 ◦ C at a rate of 1.60% min−1 for [PyFor] regener-
icant changes in the glass transition temperatures of kraft lignin
ated lignin. This could be due to the differences in the C C linkages
and regenerated lignin are found. The small difference in Tg of kraft
as reported by Tejado et al. (2007). The results of the TGA and
lignin and [PyFor] regenerated lignin might be due to the change in
DTGA curves, indicates that the lignin obtained after dissolution
the decomposition temperature, intermolecular hydrogen bonding
and regeneration process with [PyFor] is thermally more stable
interactions and molecular weight distribution of the regenerated
than that of the kraft lignin.
lignin samples (Vasile et al., 2006).

3.4.5. Differential scanning calorimetry


Recently lignin based carbon fibre composites are attaining
much attention, and understanding the behavior of lignin during 3.5. Effect of solvent on cellulose solubility
heat treatment is extremely important for making carbon fibre
(Kadla and Kubo, 2004). Differential scanning calorimetry (DSC) The solubility of cellulose in [PyFor] and its precursors are com-
indicates the possible material softening phenomenon of the lignin pared in Fig. 13, which shows that the solubility of cellulose is very
to be used as a polymer blend. The glass transition temperatures low and almost considered as negligible in [PyFor] (i.e, <1% (w/w)),
“Tg ” of kraft lignin and regenerated lignin were determined using which is considered as a desirable as well as an expected property of
the DSC thermograms (Fig. 12)), and the glass transition tempera- a solvent for selective dissolution of lignin, from the lignocellulosic
tures are 93.72 ◦ C and 98.44 ◦ C for [PyFor] regenerated lignin and biomass.
292 T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293

recycling was found to be more than 98%, which indicated that


[PyFor] as a potentially good recyclable solvent. The solvent thus
recovered was used for three cycles of lignin dissolution. The dis-
solution efficiencies are 97.2%, 85.6% and 80.2% for recycling cycles
1, 2 and 3 respectively (Fig. 15). The reduction in dissolution effi-
ciencies are due to the water content of the solvent i.e., 2%, 4.7%
and 7.8% for 1, 2 and 3 cycles respectively. During the recycling
experiments the water content was not reduced to ppm level. The
dissolution efficiency could be improved by reducing the water
content to appropriate level during the recycling operations.

3.8. Characterization of the recycled solvent

During the regeneration and recycling process, the solvent


purity and stability after each cycle were determined using 1 H NMR.
The 1 H NMR spectra for [PyFor] was recorded for first and second
recycle (Fig. 16). It was observed that the spectra in case of first and
second recycle are comparable to the pure [PyFor] which indicates
that no side products are formed during the dissolution process and
the molecular structure of the [PyFor] was not disturbed (Glas et al.,
2015 Mainka et al., 2015).

4. Conclusions

Pyridinium formate [PyFor] showed a high capacity for the


dissolution of kraft lignin (70% w/w) at a relatively lower tempera-
ture (75 ◦ C). The present study showed that increasing the alkyl
chain length and hydrogen bond basicity resulted in decreased
lignin solubility. The characteristics of the recovered lignin showed
a satisfactory comparison with the pure kraft lignin. The recov-
ered lignin showed much lower polydispersity and the average
molecular weight was reduced from 4119 g/mol to 1249 g/mol. Fur-
thermore, the regenerated lignin with increased thermal stability
was obtained and the glass transition temperature of regenerated
lignin was not affected by the dissolution and regeneration process.
The results of the experiments on regeneration and recyclabil-
Fig. 16. Comparison of 1 H NMR spectra of (A) pure [PyFor], (B) [PyFor] after first ity of [PyFor] demonstrated that the present dissolution process
recycle and (C) [PyFor] after second recycle.
using Protic Ionic liquid could lead to a diversification towards
the production of high value added end products from lignin such
3.6. Effect of lignin contents in the presence of MCC on lignin as; phenolic compounds, binders, carbon fibers, motor fuel, lignin
recovery based polymers and sorbents. This could lead to the development
of lignin-based bio refinery and significantly enhance the pretreat-
Effect of lignin contents in lignocellulose is of great importance, ment techniques for lignocellulose processing. Based on the present
as reported earlier, the lignin content in normal hardwood and soft- lead, further research in this direction, has been instigated to prove
wood varies from 20% to 30% and 26% to 32% respectively. The the application potential and efficiency of [PyFor] solvent for the
structure and composition of lignin varies from source to source extraction of lignin from real lignocellulosic biomass.
and is highly dependent on atmospheric conditions (Horvath,
2006). Mixtures of pure kraft lignin and microcrystalline cellulose Acknowledgments
(MCC) with different compositions (10–42% (w/w)) were prepared
and same dissolution procedures were repeated for all sets and the The financial support in the form of GA (Tazien Rashid) by Uni-
results are shown in Fig. 14. It was observed that at low lignin con- versiti Teknologi PETRONAS Malaysia is highly acknowledged. The
tents it was easier to dissolve and recover lignin from the solution, authors thank CORIL, Universiti Teknologi PETRONAS for providing
but for the case of higher lignin contents it was little difficult to the instrumentation facilities for analysis.
recover the components which can be done eventually but needs
more solvent or more number of stages. The present experiments References
were conducted at single stage only.
Achinivu, E.C., Howard, R.M., Li, G., Gracz, H., Henderson, W.A., 2014. Lignin
extraction from biomass with protic ionic liquids. Green Chem. 16, 1114–1119.
3.7. Effect of solvent recycling Alriols, M.G., Tejado, A., Blanco, M., Mondragon, I., Labidi, J., 2009. Agricultural
palm oil tree residues as raw material for cellulose: lignin and hemicelluloses
production by ethylene glycol pulping process. Chem. Eng. J. 148, 106–114.
The most important consideration to achieve economical and
Baptista, C., Robert, D., Duarte, A., 2008. Relationship between lignin structure and
environmentally friendly biomass processing is the ease of separa- delignification degree in Pinus pinaster kraft pulps. Bioresour. Technol. 99,
tion of lignin from lignocellulose and the recyclability of the solvent 2349–2356.
without losing its extraction efficiency. To demonstrate this, the Baumberger, S., Abaecherli, A., Fasching, M., Gellerstedt, G., Gosselink, R., Hortling,
B., Li, J., Saake, B., de Jong, E., 2007. Molar mass determination of lignins by
recovered solvent was subjected to rotary vacuum evaporation for size-exclusion chromatography: towards standardisation of the method.
6 h at 80 ◦ C and 30–kpa. The recovered mass of the [PyFor] after Holzforschung 61, 459–468.
T. Rashid et al. / Industrial Crops and Products 84 (2016) 284–293 293

Belieres, J.-P., Angell, C.A., 2007. Protic ionic liquids: preparation, characterization, Muhammad, N., Man, Z., Bustam, M.A., Mutalib, M.I.A., Rafiq, S., 2013.
and proton free energy level representation. J. Phys. Chem. B 111, 4926–4937. Investigations of novel nitrile-based ionic liquids as pre-treatment solvent for
Chaubey, A., Pandeya, S., 2011. Pyridine: a versatile nucleuse in pharmaceutical extraction of lignin from bamboo biomass. J. Ind. Eng. Chem. 19, 207–214.
field. Asian J. Pharm. Clin. Res. 4, 5–8. Ni, Y., Hu, Q., 1995. Alcell lignin solubility in ethanol–water mixtures. J. Appl.
da Costa Lopes, A.M., João, K.G., Rubik, D.F., Bogel-Łukasik, E., Duarte, L.C., Andreaus, Polym. Sci. 57, 1441–1446.
J., Bogel-Łukasik, R., 2013a. Pre-treatment of lignocellulosic biomass using Pinkert, A., Goeke, D.F., Marsh, K.N., Pang, S., 2011. Extracting wood lignin without
ionic liquids: wheat straw fractionation. Bioresour. Technol. 142, 198–208. dissolving or degrading cellulose: investigations on the use of food
da Costa Lopes, A.M., João, K.G., Morais, A.R.C., Bogel-Łukasik, E., Bogel-Łukasik, R., additive-derived ionic liquids. Green Chem. 13, 3124–3136.
2013b. Novel pretreatment and fractionation method for lignocellulosic Pu, Y., Jiang, N., Ragauskas, A.J., 2007. Ionic liquid as a green solvent for lignin. J.
biomass using ionic liquids. RSC Adv. 3, 16040–16050. Wood Chem. Technol. 27, 23–33.
Erdocia, X., Prado, R., Corcuera, M.A., Labidi, J., 2014. Effect of different organosolv Remsing, R.C., Hernandez, G., Swatloski, R.P., Massefski, W.W., Rogers, R.D., Moyna,
treatments on the structure and properties of olive tree pruning lignin. J. Ind. G., 2008. Solvation of carbohydrates in N,N -dialkylimidazolium ionic liquids: a
Eng. Chem. 20, 1103–1108. multinuclear NMR spectroscopy study. J. Phys. Chem. B 112, 11071–11078.
Gellerstedt, G., Henriksson, G., 2008. Lignins: major sources, structure, and Rinaldi, R., 2011. Instantaneous dissolution of cellulose in organic electrolyte
properties. In: Gandini, M.N.B. (Ed.), Monomers Polymers and Composites from solutions. Chem. Commun. 47, 511–513.
Renewable Resources. Elsevier, Amsterdam, pp. 201–224. Santos, P.S., Erdocia, X., Gatto, D.A., Labidi, J., 2014. Characterisation of kraft lignin
George, A., Brandt, A., Tran, K., Zahari, S.M.N.S., Klein-Marcuschamer, D., Sun, N., separated by gradient acid precipitation. Ind. Crops Prod. 55, 149–154.
Sathitsuksanoh, N., Shi, J., Stavila, V., Parthasarathi, R., 2015. Design of low-cost Scriven, E.F.V., Murugan, R., 2000. Pyridine and pyridine derivatives. In:
ionic liquids for lignocellulosic biomass pretreatment. Green Chem. 17, Kirk-Othmer Encyclopedia of Chemical Technology, fourth ed. John Wiley &
1728–1734. Sons, Inc.
Glas, D., Van Doorslaer, C., Depuydt, D., Liebner, F., Rosenau, T., Binnemans, K., De Shimizu, S., Watanabe, N., Kataoka, T., Shoji, T., Abe, N., Morishita, S., Ichimura, H.,
Vos, D.E., 2015. Lignin solubility in non-imidazolium ionic liquids. J. Chem. 2000. Pyridine and pyridine derivatives. In: Ullmann’s Encyclopedia of
Technol. Biotechnol. 90, 21–27. Industrial Chemistry. Wiley-VCH Verlag GmbH & Co.
Hart, W.E.S., Harper, J.B., Aldous, L., 2015. The effect of changing the components of Sun, S., Li, M., Yuan, T., Xu, F., Sun, R., 2013. Effect of ionic liquid/organic solvent
an ionic liquid upon the solubility of lignin. Green Chem. 17, 214–218. pretreatment on the enzymatic hydrolysis of corncob for bioethanol
Haviv, F., DeNet, R.W., Michaels, R.J., Ratajczyk, J.D., Carter, G.W., Young, P.R., 1983. production. Part 1: structural characterization of the lignins. Ind. Crops Prod.
2-[(Phenylthio) methyl] pyridine derivatives: new anti-inflammatory agents. J. 43, 570–577.
Med. Chem. 26, 218–222. Sun, Y., Cheng, J., 2002. Hydrolysis of lignocellulosic materials for ethanol
Horvath, A.L., 2006. Solubility of structurally complicated materials: I. Wood. J. production: a review. Bioresour. Technol. 83, 1–11.
Phys. Chem. Ref. Data 35, 77–92. Sundquist, J., 1988. Problems of Non-conventional Pulping Processes in the Light of
Ji, W., Ding, Z., Liu, J., Song, Q., Xia, X., Gao, H., Wang, H., Gu, W., 2012. Mechanism of Peroxyformic Acid Cooking Experiments. The Finnish Pulp and Paper Research
lignin dissolution and regeneration in ionic liquid. Energy Fuels 26, 6393–6403. Institute, Helsinki.
Kadla, J.F., Kubo, S., 2004. Lignin-based polymer blends: analysis of intermolecular Swatloski, R.P., Spear, S.K., Holbrey, J.D., Rogers, R.D., 2002. Dissolution of cellose
interactions in lignin—synthetic polymer blends. Compos. Part A Appl. Sci. with ionic liquids. J. Am. Chem. Soc. 124, 4974–4975.
Manuf. 35, 395–400. Tejado, A., Peña, C., Labidi, J., Echeverria, J.M., Mondragon, I., 2007.
King, A.W.T., Zoia, L., Filpponen, I., Olszewska, A., Xie, H., Kilpeläinen, I., Physico-chemical characterization of lignins from different sources for use in
Argyropoulos, D.S., 2009. In situ determination of lignin phenolics and wood phenol—formaldehyde resin synthesis. Bioresour. Technol. 98, 1655–1663.
solubility in imidazolium chlorides using 31P NMR. J. Agric. Food Chem. 57, Vasile, C., Gosselink, R., Quintus, P., Koukios, E., Koullas, D., Avgerinos, E., Abacherli,
8236–8243. D., 2006. Analytical methods for lignin characterization. I. Thermogravimetry.
Kubo, S., Kadla, J.F., 2005. Hydrogen bonding in lignin: a fourier transform infrared Cellul. Chem. Technol. 40, 421–429.
model compound study. Biomacromolecules 6, 2815–2821. Vila, C., Santos, V., Parajo, J.C., 2003. Recovery of lignin and furfural from acetic
Lateef, H., Grimes, S., Kewcharoenwong, P., Feinberg, B., 2009. Separation and acid—water—HCl pulping liquors. Bioresour. Technol. 90, 339–344.
recovery of cellulose and lignin using ionic liquids: a process for recovery from Vitz, J., Erdmenger, T., Haensch, C., Schubert, U.S., 2009. Extended dissolution
paper-based waste. J. Chem. Technol. Biotechnol. 84, 1818–1827. studies of cellulose in imidazolium based ionic liquids. Green Chem. 11,
Lee, S., Doherty, T., Linhardt, R., Dordick, J., 2009. Ionic liquid-mediated selective 417–424.
extraction of lignin from wood leading to enhanced enzymatic cellulose Wang, Q., Chen, K., Li, J., Yang, G., Liu, S., Xu, J., 2011. The solubility of lignin from
hydrolysis. Biotechnol. Bioeng. 102, 1368–1376. bagasse in 1:4-butanediol/water system. Bioresources 6, 3034–3043.
Li, M.-F., Sun, S.-N., Xu, F., Sun, R.-C., 2012. Formic acid based organosolv pulping of Zakrzewska, M.E., Bogel-Łukasik, E., Bogel-Łukasik, R., 2010. Solubility of
bamboo (Phyllostachys acuta): comparative characterization of the dissolved carbohydrates in ionic liquids. Energy Fuels 24, 737–745.
lignins with milled wood lignin. Chem. Eng. J. 179, 80–89. Zhang, J., Deng, H., Lin, L., Sun, Y., Pan, C., Liu, S., 2010. Isolation and
Maga, J.A., 1981. Pyridines in foods. J. Agric. Food Chem. 29, 895–898. characterization of wheat straw lignin with a formic acid process. Bioresour.
Mai, N.L., Ha, S.H., Koo, Y.-M., 2014. Efficient pretreatment of lignocellulose in ionic Technol. 101, 2311–2316.
liquids/co-solvent for enzymatic hydrolysis enhancement into fermentable Zhang, Y.-H.P., 2008. Reviving the carbohydrate economy via multi-product
sugars. Process Biochem. 49, 1144–1151. lignocellulose biorefineries. J. Ind. Microbiol. Biotechnol. 35, 367–375.
Mainka, H., Tager, O., Körner, E., Hilfert, L., Busse, S., Edelmann, F.T., Herrmann, A.S., Zhou, S., Liu, L., Wang, B., Xu, F., Sun, R., 2012. Microwave-enhanced extraction of
2015. Lignin—an alternative precursor for sustainable and cost-effective lignin from birch in formic acid: structural characterization and antioxidant
automotive carbon fiber. J. Mater. Res.Technol. 4, 283–296. activity study. Process Biochem. 47, 1799–1806.
Maki Arvela, P., Anugwom, I., Virtanen, P., Sjoholm, R., Mikkola, J.P., 2010.
Dissolution of lignocellulosic materials and its constituents using ionic
liquids—a review. Ind. Crops Prod. 32, 175–201.

Potrebbero piacerti anche