Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Accepted Manuscript

Coconut coir pith lignin: A physicochemical and thermal


characterization

L. Asoka Panamgama, P.R.U.S.K. Peramune

PII: S0141-8130(17)34191-0
DOI: doi:10.1016/j.ijbiomac.2018.03.012
Reference: BIOMAC 9234
To appear in:
Received date: 26 October 2017
Revised date: 22 February 2018
Accepted date: 2 March 2018

Please cite this article as: L. Asoka Panamgama, P.R.U.S.K. Peramune , Coconut coir pith
lignin: A physicochemical and thermal characterization. The address for the corresponding
author was captured as affiliation for all authors. Please check if appropriate.
Biomac(2017), doi:10.1016/j.ijbiomac.2018.03.012

This is a PDF file of an unedited manuscript that has been accepted for publication. As
a service to our customers we are providing this early version of the manuscript. The
manuscript will undergo copyediting, typesetting, and review of the resulting proof before
it is published in its final form. Please note that during the production process errors may
be discovered which could affect the content, and all legal disclaimers that apply to the
journal pertain.
ACCEPTED MANUSCRIPT

Coconut Coir Pith Lignin: A Physicochemical and Thermal Characterization


L. Asoka Panamgama* and P. R. U. S. K. Peramune
Department of Chemistry, Faculty of Science, University of Ruhuna, Matara, Sri Lanka
asoka@chem.ruh.ac.lk
umeshawinzz@gmail.com

*Correspondence author: L. Asoka Panamgama

PT
E-mail: asoka@chem.ruh.ac.lk
Telephone: +(94)-41-2222681+(94)-41-2222682
Fax: (+94)-41-2227001

RI
SC
NU
MA
D
P TE
CE
AC

1
ACCEPTED MANUSCRIPT

Graphical Abstract

PT
RI
SC
NU
MA

Highlights
D

 Three extraction protocols (viz., alkaline, organosolv and PEG) incorporating two energy
TE

sources (viz., thermal and microwave) were applied to coconut coir pith feedstock for
lignin extraction.
P
CE

 Influential parameters on lignin extraction protocols were obtained for various energy
supply sources.
AC

 Physicochemical and thermal characteristics of different lignin structures were identified


by FTIR, TGA and wet chemical analysis.

Abstract
The structural and thermal features of coconut coir pith lignin, isolated by three different
extraction protocols incorporating two different energy supply sources, were characterized by
different analytical tools. The three different chemical extraction protocols were alkaline - 7.5
% (w/v) NaOH, organosolv - 85 % (v/v) formic and acetic acids at 7:3 (v/v) ratio and

2
ACCEPTED MANUSCRIPT

polyethylene glycol (PEG): water ratio at 80:20 %wt. The two sources of energy were
thermal or microwave. Raw lignins were modified by epichlorohydrin to enhance reactivity,
and the characteristics of raw and modified lignins were comparatively analysed. Using the
thermal energy source, the alkaline and organosolv processes obtained the highest and lowest
lignin yields of 26.4 ±1.5 %wt and 3.4 ±0.2 %wt, respectively, as shown by wet chemical
analysis. Specific functional group analysis by Fourier transform infrared spectra (FTIR)
revealed that significantly different amounts of hydroxyl and carbonyl groups exist in

PT
alkaline, organosolv and PEG lignins. Thermogravimetric analysis (TGA) illustrated that the
lowest degradation onset temperature was recorded for organosolv lignin, and the overall
order was organosolv < alkaline ≤ PEG (microwave) < PEG (thermal). Irrespective of the

RI
extraction protocol, microwave energy provided the highest %wt loss rate, indicating the

SC
lowest thermal stability. The derivative temperature difference profiles from the microwave
and thermal heating sources for different extraction protocols are discussed in detail. These
NU
findings show that lignin extraction from coir pith can be performed efficiently with several
protocols and that those methods offer practical value to industry.
MA

Keywords Coconut coir pith, Lignocellulosic waste, Microwave, FTIR spectroscopy,


Thermal analysis
D
TE

1. Introduction
Coconut (Cocos nucifera L.), a member of the Arecaceae family (palm family), is one of the
major plantation crops, and it accounts for approximately 12% of all agricultural produce in
P

Sri Lanka. Total land area under coconut cultivation is 3.95 x 105 hectares, and approximately
CE

2.5 x 103 million nuts are produced each year [1]. Sri Lankan desiccated coconut, produced
from the kernel, is highly popular worldwide for its characteristic taste and distinguishable
AC

white colour, and it holds a high ranking in the world export market. Coconut coir is
produced from the tissues surrounding the seed of the coconut palm (i.e., the coconut husk),
and a technique called the drum system is used to extract the coir fibre, resulting in long pure
fibre. Coir fibres are odourless, lightweight, thick, strong and resist abrasion. As such, these
fibres are used to produce various industrially useful articles, such as doormats, rugs, sack
cloth, brushes, mattresses, insulation panels, and packaging. [2]. Furthermore, Sri Lanka
holds the number one position worldwide for the export of brown coconut fibre. Coconut coir
pith is a light brown coloured, spongy, lignocellulosic waste biomaterial that is generated in
huge volumes during the processing of coir fibre from coconut husks. Coir pith falls from the

3
ACCEPTED MANUSCRIPT

thick mesocarp of the coconut fruit when it is shredded during coir fibre processing, and it
has a porous structure that typically consists of particles ranging from 0.2 to 2.0 mm (75 % to
90 %) in size [3]. Generally, one unit-mass of coir fibre generates two unit-masses of coconut
coir pith, which remains in huge quantities as a waste product with low industrial value that is
available plentifully in Sri Lanka [2]. Characteristics of coir pith are its hydrophilic nature,
absence of pathogens and renewability with no ecological drawbacks to its use due to its slow
natural decomposition rate. Coir pith consists primarily of lignin cellulose, hemicelluloses

PT
and extractives, such as pectins and tannins, and the %wt content of lignin is high [3]. These
constituents primarily contain carboxyl, hydroxyl and ether groups, which makes coir pith a
useful adsorbent or a natural ion exchanger. Generally, coir pith can be incinerated or

RI
dumped without any limitation [4]. The density of coir pith is typically > 0.2 g cm-3, and it

SC
can be highly compressed (by a factor of 8 to 10) in order to make coir bricks (Fig. 1), which
would facilitate its transportation and exportation. Coir pith is also used in the preparation of
NU
soilless growing media for the cultivation of some ornamental plants and crops [5].
Alternative, widespread usage of this lignocellulosic waste biomaterial has succeeded due to
its high-water absorption and retention characteristics. Other uses of coir pith and its
MA

composites have also been documented in the literature [6-8].


D

Lignocellulosic waste is receiving widespread attention as a promising, renewable,


TE

abundantly available, low-cost material [9]. In plant materials, lignin is one of the major
components that is present in variable proportions. In a majority of the industries that address
the treatment of lignocellulosic materials, especially in pulp and paper production, lignin is
P

one of the primary by-products [10]. The low reactivity of lignin is a drawback, and it is
CE

currently regarded by most as a waste material [11]. The reduction of synthetic adhesive
usage in wood composite production by the development of lignocellulosic waste materials is
AC

a challenge [12]. Lignin, possessing complex chemical structures, poses a potential barrier to
many applications seeking to replace expensive petrochemical-based materials [13]. The
main challenge to the widespread usage of lignin is understanding its complex structure and
physicochemical properties [14-18].

Lignin is a complex phenolic biological macromolecule, built by oxidative coupling of three


major C6–C3 (phenyl propanoid) units, namely, syringyl (S), guaiacyl (G), and p-coumaryl
(H) units, which form randomized three-dimensional network structures inside cell walls
[19]. The major inter-unit linkages in lignin are of the aryl–aryl and ether types. It is reported

4
ACCEPTED MANUSCRIPT

that there are twenty different types of C-C and C-O bonds present in lignin itself. The bonds
include -OCH3, -C=O (OH) and -C=O. There are also a variety of ether linkages, such as α-
O-4, β-O-4 and 4-O-5 with numerous C–C bonds, such as β-β, β-1, β-5 and 5-5 [20, 21]. The
typical characteristics of isolated lignins depends on several types of factors, such as variety,
geographic area, location, soil type and season. The nature of the extraction protocol also
plays a dominant role. Structurally different lignins have been extracted by changing the
extraction protocol [9, 15].

PT
In thermal heating, heat transfer to the reaction mixture occurs homogeneously and slowly;
thus, all points within the reaction medium reach approximately the same temperature [22].

RI
Microwaves are electromagnetic in nature and possess electric and magnetic fields that are

SC
perpendicular to each other. In microwave usage, heat generation occurs via two
simultaneous mechanisms, viz., dipolar rotation and ionic conduction, which liberates
NU
thermal energy into the medium, thereby resulting in a very rapid and homogeneous increase
in temperature in the reactor [23]. Both the solvent and the solid matrix strongly affect the
nature of microwave energy interaction efficiency [22]. Solvents with larger dielectric
MA

constants, which possess permanent dipole moments, facilitate microwave energy absorption
and are optimal for use [24]. Microwave energy assisted wood and bark liquefaction and
D

lignin extraction processes have been studied by a number of researchers [23, 25-27]. One of
TE

the advantages of microwave assisted heating is the disruption of weak hydrogen bonds,
which promotes the dipole rotation of the molecules. Several physical parameters of solvents
(e.g., dielectric constants, dipole moments and dissociation constants) that were used in the
P

present study are shown in Table S1, supplementary materials. If coconut coir pith can be
CE

utilized as a renewable bio-resource to extract lignin, it would provide sizeable economic


advantages. Therefore, the objectives of this research were (i) calculation of the yields of coir
AC

pith lignin by different extraction protocols; (ii) modification and characterization of coir pith
lignin by epoxidation such that it could be used in preparation of polymers and bio-
composites in a feasibility study; and (iii) comparison of the thermal and physicochemical
properties of lignins obtained by the three different protocols with variation in energy supply
sources, i.e., microwave assisted extraction versus conventional thermal energy assisted
extraction.

2. Materials and Methods


2.1 Materials

5
ACCEPTED MANUSCRIPT

Coconut coir pith, as compressed blocks with dimensions of 20 x 10 x 5 (± 0.5) cm3 and
weights in the range of 550 - 650 g, were generously supplied by Hayleys Fibre Plc.
(Colombo, Sri Lanka) (Fig. 1). Coir pith blocks were crushed into powder in a hammer mill,
sieved by a 60-mesh size filter to separate fine particles, and oven dried at 60°C for ~ 12 h
before use. The initial moisture content of the coir pith was 8 %wt and this was reduced to 2
± 1 %wt. All chemicals used were of analytical grade, including NaOH pellets (≥ 99.5 %),
acetic acid (99.8%), formic acid (99.5 %), concentrated sulfuric acid (≥ 95 %), Polyethylene

PT
glycol -400, epichlorohydrin (99 %), and 1,4-Dioxane (99.5 %), except for KBr, which was
spectroscopy grade. All of the chemicals were purchased from Sigma Aldrich Co. (St. Louis,
MO, USA) and were used without any further purification. Distilled water was used for all

RI
synthesis and modification stages.

SC
2.2 Thermal energy assisted extraction of lignin by different protocols
NU
Coir pith lignin extraction was performed with conventional glassware using an oil bath and
reflux setup. To maintain homogeneity in the reaction, the reactor vessel was manually stirred
with a wire spring every 30 sec. All lignin extraction protocols were repeated under the same
MA

reaction conditions, and the %wt of the lignin yields were calculated for each protocol. The
values reported in Table 1 are representative of the mean values of at least three replicate
D

measurements.
TE

2.2.1 Alkaline protocol


A slightly modified alkaline protocol was followed according to A. Tejado et al. [28]. In
P

brief, a mixture of 7.5% (w/v) NaOH and coir pith in solvent at a 1:10 (w/v) ratio was
CE

maintained. These materials were refluxed in a 3-neck round bottom flask at 90 ± 2 °C for 1.5
h. After refluxing, two components were observed, a solid residue and a liquid. This liquid
AC

component is typically called “black liquor” due to its characteristic dark colour. The pH of
this black liquor was approximately 12. While the black liquor was still warm, the solid
residue was separated from the black liquor by filtering it with a silk cloth by manually
squeezing it, and then it was cooled to the ambient temperature of 27 °C by dipping it in a
cold water bath at 10 °C. The separated black liquor was acidified with 2 M H2SO4 until the
pH decreased to near 2 and was then adjusted to exactly pH = 2 by adding 0.5 M H2SO4 drop-
wise with a pipette while stirring, which facilitated alkaline lignin precipitation and
sedimentation. During acidification, the processed of crude colloid lignin precipitation and
sedimentation were very rapid and easily distinguishable, as the blackish colour changed to a

6
ACCEPTED MANUSCRIPT

light brownish-yellow. The precipitated lignin was later separated in a Buckner funnel and
washed with distilled water until it reached pH neutrality and oven dried at 50 ± 2 °C for 48
h. Dry lignin was powdered using a mortar and pestle until it could pass through a 60-mesh
sieve (250 µm) and was stored in zipped plastic bags for further use.

2.2.2 Organosolv protocol


The organosolv protocol used to extract lignin from coir pith followed a slightly modified

PT
process from that described in the literature [29-31]. In this process, 85 % (v/v) acetic acid
and 85 % (v/v) formic acid was used in a 7:3 (v/v) ratio, and coconut coir pith was added to
solvent at a 1:8 (w/v) ratio. The mixture was refluxed in a 3-neck round bottom flask for 2 h

RI
at 98 ± 2 °C and allowed to cool to the ambient temperature of 27 °C. Unreacted solid residue

SC
was separated from the black liquor by filtering with a silk cloth and by manually squeezing
it. The solid residue was washed with a small volume of 80 % (v/v) (~ 2 x 10 mL) formic
NU
acid and with a small volume (~10 mL) of distilled water until the rinse became clear. All of
the washings were collected, and the black liquor was diluted four times with distilled water,
thoroughly stirred and centrifuged. This process facilitated colloidal lignin sedimentation.
MA

Lignin recovery procedures followed and the subsequent steps were the same as for the
alkaline process.
D
TE

2.2.3 Polyethylene glycol (PEG-400) assisted extraction protocol


The protocol followed was a slightly modified procedure from what has been described
elsewhere [32]. In this instance, the solvent was made up of PEG mixed with water at an
P

80:20 %wt ratio. Next, coconut coir pith was mixed with this solvent at a 1:4 (w/w) ratio and
CE

used as a de-lignifying medium with the addition of 1 %wt concentrated H2SO4, based on
coir pith. The mixture was refluxed in a 3-neck round bottom flask at 160 ± 2 °C for 2 h and
AC

allowed to cool to the ambient temperature of 27 °C. After cooling, the pulp residue was
removed by filtering, using a silk cloth and manually squeezing. The solid residue was later
washed with a small volume of 1,4-dioxane, and the mixture was well-stirred. The 1,4-
dioxane in the black liquor was recovered by rotary evaporation. The black liquor was diluted
10 times with distilled water, thoroughly stirred, centrifuged and allowed to settle for 1 h for
lignin recovery. The subsequent procedures followed were the same as for the alkaline
protocol.

2.3 Microwave-assisted lignin extraction protocols

7
ACCEPTED MANUSCRIPT

A household microwave oven (LG, model MS-2043DB) with a rotating turntable was
purchased from Abans Ltd., (Colombo, Sri Lanka). The operating frequency and power
output were 2450 MHz and 900 W, respectively. Typical extraction recipe formulations were
the same as with the thermal energy assisted extraction protocol. The materials and reagents
were placed in a Teflon tube, screw-capped to seal, placed in the microwave oven, and
allowed to react for the required times (5 - 20 min.). After the reaction, the Teflon tube was
quenched to an ambient temperature of 27 °C before opening for lignin recovery by

PT
immersion in a cold water bath at 10 °C. Extraction procedures and the subsequent steps
followed were the same as with the thermal energy assisted extraction processes. Note that no
weight changes were observed in the sealed tubes before and after heating.

RI
SC
2.4 Characterization
TAPPI standards were used to determine the chemical composition of the coconut coir pith
NU
raw material. Moisture content was determined gravimetrically after drying the sample at 105
±3 °C for 24 h (TAPPI T264 cm-97), and ash content was determined via incineration at 575
°C (T211 om-93). The results of these analyses are shown in Table 1. Lignins obtained by the
MA

different extraction protocols were not subjected to any additional purification steps. The
alkaline lignin and PEG lignin obtained by thermally assisted extraction protocols were
D

chemically modified to increase their reactivity by epoxidation with epichlorohydrin. All of


TE

the raw and chemically modified lignins obtained from the different extraction protocols were
characterized by different techniques to analyse their structural integrity.
P

2.4.1 Fourier transform infrared spectroscopy (FTIR)


CE

All of the lignin samples were ground with spectroscopy grade KBr (Sigma-Aldrich) -based
on a 1:99 %wt lignin to KBr ratio- pressed into pellets, and analysed on a Thermo Scientific
AC

Nicolet iS10 FTIR Spectrometer (Thermo Nicolet Corp., Madison, WI, USA) in transmission
mode at the room temperature of 27 °C. Spectra were obtained by taking the mean of 16
scans at 4 cm−1 resolution over the 400 – 4000 cm-1 range. All spectral processing
(calibration, background subtraction, and automatic baseline correction) was achieved using
Omnic 7.0 software. Spectral intensity data comparisons were performed based on peak
height measurements. The distance from the baseline to the peak top was used as the peak
height. The aromatic (phenyl) C=C skeletal vibration peak appearing at 1510 cm-1 was
subjected to no changes; therefore, it was used to normalize spectral peaks by setting its
intensity to 1 in each spectrum. An FTIR spectral peak height data, based semi-quantitative

8
ACCEPTED MANUSCRIPT

analysis, was developed in which the ratio Ix/I1510 represented the quotient between the
transmittance of selected peaks with a reference peak at 1510 cm−1.

2.4.2 Thermal analysis


Thermal analysis was performed using a simultaneous thermal analyser SDT Q600
(TGA/DSC) V20.9 Build 20 (TA Instrument, Waters LLC, USA). TA Universal Analysis
2000 software was used for the analysis of thermal data. The instrument was calibrated

PT
against Indium and Zinc standards. Precision of the instrument varies over a temperature
range from 27 (ambient) to 400 °C by ± 0.5 °C and from 400 to 800 °C by ±1 °C according to
the calibration procedure. A flat α-Al2O3 empty crucible was used as the reference material.

RI
The sample weights varied between 5 and 8 (± 0.25) mg. All samples were analysed under a

SC
nitrogen atmosphere of 100.00 mL min-1 at a constant heating rate of 10 °C min-1. Analyses
were performed from a temperature of 27 °C (ambient) up to 800 °C. This instrument is
NU
capable of simultaneously analysing eight different thermal parameters for a given material.
All of the parameters acquired by this instrument during thermal analysis are given in Table
S2 (supplementary materials). Moreover, the measurement of derivative temperature
MA

differences (°C/°C) with variation in temperature is another feature of this instrument (see
Fig. S1 (a) - (e), supplementary materials).
D
TE

2.4.3 Statistical analysis


The data are expressed as the mean of three or six independent experiments ± standard error
(SE). One-way variance analysis (ANOVA) was performed to test the significance of the
P

observed differences using the statistical analysis package SPSS 12.0 (SPSS Inc., Chicago,
CE

IL, USA), and the results were compared using Duncan’s multiple range tests at a 95 %
confidence level.
AC

3. Results and Discussion


3.1 Lignin yield and ash content analysis
Coconut coir pith is an abundantly available, underutilized waste material in Sri Lanka.
Klason lignin content of raw coir pith was observed to be notably high [5, 6]. In the first part
of this research, inexpensive chemicals were used for the extraction of lignin, and the
physicochemical properties of coir pith lignins were elucidated. Interestingly, no clear trend
could be observed in the correlation between %wt lignin yields, obtained with different

9
ACCEPTED MANUSCRIPT

extraction protocols, and the corresponding physical parameters of solvent data (Table S1,
supplementary materials).

In determining the ash content, the powdered lignin samples were held at 575 °C until they
reached constant weights. It is reported that there are two types of ash in lignocellulosic
materials viz., non-extractable and extractable. Inorganic materials that are bound to the
structure of lignocelluloses comprise the non-extractable ash. Extractable ash is the inorganic

PT
matter that adheres to the biomass and can be removed by washing with water; it includes
windborne soil, sand particles and various metal oxides, primarily of sodium that is added in
the extraction process and minor quantities of iron and potassium [33]. The physicochemical

RI
data (mean %wt lignin yield and ash content) of lignins obtained by both thermal and

SC
microwave assisted extraction conditions are given in Table 1 with different protocols
according to their original coir pith weights. It can be observed that alkaline rice straw lignin
NU
has the highest ash content of 11.1±1.2 %wt. Alkaline coir pith lignin extraction protocols
have reported higher ash contents (~ 10 %wt) than other processes irrespective to the mode of
energy supply, but the numerical values were nearly the same within experimental errors. The
MA

organosolv protocol produced moderate ash content values (~ 6 %wt) For coir pith lignin, the
ash content present can be given in ascending order as follows: alkaline > organosolv > PEG.
D
TE

Generally, microwave process has drastically reduced reaction times with the removal of
cumbersome glassware and other procedures while having significantly high yields [22, 23].
Therefore, usage of microwave energy in lignin extraction has significant advantages with a
P

prominent gain in high %wt yields [24 - 26]. The alkaline thermal energy assisted extraction
CE

process gave the highest lignin yield of 26.4±1.5 %wt based on biomass (Table 1). In this
process, an instant, 10-unit change occurred in the pH value of the black liquor. The change
AC

from a basic pH of 12 to an acidic pH of 2 was very distinct and rapid, and it facilitated lignin
colloidal formation followed by precipitation. In the microwave alkaline process, lignin yield
increased up to 32.4 ±1.6 %wt, whereas in the organosolv process, no such distinct pH
change occurred; only a dilution followed by centrifugation was observed. Therefore, the
organosolv process seems to less effectively facilitate micro-colloidal lignin coagulation
followed by sedimentation. In the organosolv process, several micro-colloidal lignin may
have washed away and been wasted and not accounted. As a result, the organosolv process
reported the lowest yield of 3.4 ±0.2 %wt (Table 1). In the microwave, the organosolv
process lignin yield has increased to 4.1 ±0.4 %wt. For PEG, thermal energy assisted lignin

10
ACCEPTED MANUSCRIPT

extraction had a yield of 6.7 ±0.6 %wt, and when the mode of energy supply was changed to
microwave, the yield increased to 7.2 ±0.8 %wt.

It has been reported that C–C bond cleavage is difficult to achieve under thermal heating but
occurs relatively easily under microwave conditions, forming more mono-phenolic
compounds [23]. This phenomenon could be the reason for the higher %wt yields acquired
under microwave heating. In addition, the comparatively low yields obtained in this study

PT
could be due to the less efficient isolation of lignin, partly by stronger linkages to
hemicelluloses [27], as well as wastage due to partial precipitation and sedimentation of
micro-colloidal lignin particles, since their small size does not facilitate their retention with

RI
the bulk lignin. Nevertheless, not considering the energy source, an approximate comparative

SC
quantitative analysis on %wt lignin yields was observed to have a clear trend, as follows:
Alkali > PEG > organosolv.
NU
3.2 FTIR spectral analysis of raw and modified lignin
Non-destructive FTIR spectroscopy can be used to serve as a qualitative indication of purity
MA

and structural similarity in lignins [34]. As lignins contain various functional groups, the
physicochemical properties of lignin samples obtained by different extraction protocols were
D

analysed by FTIR spectroscopy. FTIR spectra of selected thermal energy assisted unmodified
lignins are shown in Fig. 2 (a) and the magnified region of 2000 to 400 cm−1 is shown in Fig.
TE

2 (b). It can be seen that all four unmodified lignins have similar absorption patterns, which
corresponds to their inherent functional groups, indicating the existence of analogous
P

structures. The identification of typical functional groups were performed according to the
CE

literature reported methods [34-37]. However, the peak intensities (heights) and positions of
specific functional groups of the lignin spectra differed slightly and this variation could be
AC

explained not only by variations in the biomass source but also by the nature of the extraction
protocol.

Peaks appearing in the range of 1600 (1618 - 1607) cm-1, 1510 (1516 - 1514) cm-1 and 1424
(1463 - 1424) cm-1 correspond to the aromatic C=C ring vibrations in phenyl-propane
structures (35, 36). All lignins have shown a broad absorption band in the 3442 - 3312 cm-1
range, which can be assigned to the -OH stretching vibration, due to the presence of either
alcohol, hydroxyl or phenolic hydroxyl groups in the aromatic and aliphatic moieties of lignin
[38]. Extraction protocols have affected the extent of hydrogen bonding in lignins, and an

11
ACCEPTED MANUSCRIPT

analysis of -OH absorption peak positions can be useful to understand the nature of hydrogen
bonding. Alkaline coir pith lignin has the least hydrogen bonds, with a peak appearing at
3450 cm-1, while organosolv coir pith lignin has the most hydrogen bonds, with the peak
being shifted to 3290 cm-1. In a comparison of rice straw lignin with PEG coir pith lignin, the
rice straw has more hydrogen bonds, and the –OH peak appears at 3407 cm-1. Incorporation
of PEG into the lignin moiety has facilitated more hydrogen bond formation, as indicated by
the stretching vibration of the –OH peak appearing at 3424 cm-1, as shown in Fig. 2 (a).

PT
The minor peaks in the range of 2917 - 2844 cm−1 can be attributed to aliphatic C–H
stretching vibrations in the -CH2- and -CH3 groups of the phenyl-propane side chains. The

RI
existence of unconjugated carbonyl ester or acetyl groups in all lignins contributes to the

SC
peaks appearing at 1732 - 1720 cm−1, whereas conjugated carbonyl ester or acetyl groups
appear in the range 1699 – 1644 cm−1. In all spectra, these two sets of peaks appear as small
NU
shoulders, indicating the presence of carbonyl ester or acetyl groups to a small extent. It has
been suggested that in native lignins, carbonyl ester or acetyl groups are present in extremely
small quantities and have usually suffered degradation during chemical extraction processes
MA

[39].
D

Miniscule peaks, appearing at 1424 cm-1 in organosolv lignin and at 1426 cm-1 in alkaline
TE

straw lignin, may indicate the presence of negligible amounts of the crystallized structure of
cellulose I. Peaks in the 1300 - 1000 cm−1 range arise from asymmetric and symmetric
vibrations due to the presence of very different functional groups, like C–O stretching in
P

phenols and ethers—C–O(H), C-O(Ar) in S and G units of lignin, as shown in Fig. 2(b) [37].
CE

A highly distinct feature observed in the alkaline straw lignin was the presence of a very
intense peak at 1246 cm-1 with a small shoulder at 1093 cm-1, which was unique and not
AC

observed in other lignins as displayed in Fig. 2(b). The bands at 1330, 1260 and 833 cm-1
indicate the presence of syringyl structural (S) units [36]. Typical bands at 1260 cm-1 and
1151 cm-1 are an indication of the breathing of associated guaiacyl (G) units with C-O
stretching. In alkaline and organosolv coir pith lignins, the 1330 cm-1 peak is very weak, and
in PEG assisted coir pith and alkaline straw lignin, this peak has shifted to 1351 cm-1 and
1384 cm-1, respectively. Peaks at approximately 1160 cm−1 indicate the presence of a p-
coumaric ester group typical of HGS type lignin. In alkaline and PEG lignin, this peak has
shifted to 1125 cm-1 and 1103 cm−1, respectively. Peaks in the range of 1035 to 1030 cm-1 are
indicative of aromatic C-H deformations, G > S, C–O in primary alcohols, and C=O

12
ACCEPTED MANUSCRIPT

stretching, and this peak has shifted to the range 1041 - 1082 cm-1 shown in Fig. 2 (b) [38].
All lignins have very weak, low-intensity peaks in the fingerprinting region of 900 to 650 cm-
1
, which are responsible for the poly-substituted phenolic structure of lignin [39]. Due to the
use of H2SO4 in the precipitation process, both alkaline rice straw and alkaline coir pith
lignins show a very intense and unique peak at 617 (616) cm-1, which is assigned to C–S
bond stretching vibration [40].

PT
Furthermore, in the liquefaction of corn stover with ethylene glycol by thermal and
microwave assisted processes, different extents of hydroxyl and carbonyl groups have been
identified, wherein the conversion of hydroxyl groups to carbonyl groups have been reported

RI
in the microwave liquefied products [39]. In this instance, the ratio of peak heights of C=C

SC
phenyl skeletal vibration peaks to carbonyl vibration peaks has been used to facilitate a
comparison of chemical changes caused in the extraction protocols followed by modification
NU
to lignin. Semi-quantitative relative peak height ratio data, obtained by different extraction
protocols, were calculated to provide the relative proportions of selected functional groups
and results, as shown in Table 3. In coir pith, the organosolv and PEG extraction protocols
MA

exhibit a high extent of carbonyl groups. Both the usage of acetic acid and formic acid and
the fragmentation process followed by chemical bonding with PEG seem to have facilitated
D

the formation of many carbonyl functional groups. Modification with epichlorohydrin has
TE

slightly lowered the carbonyl ratio in PEG lignin, whereas alkaline lignin modification has
increased the carbonyl ratio (Table 3).
P

3.3 Modification of lignins by epoxidation with epichlorohydrin


CE

When lignins are composed of guaiacyl (G) units, modification with crosslinking agents is
feasible, due to the presence of free ortho-position phenolic nuclei (C5). When lignins are
AC

composed of S units, both the C3 and C5 positions are blocked with -OCH3 groups, which
results in low reactivity with crosslinking agents. In Figs. 3(a) and 3 (b), the FTIR spectra of
epichlorohydrin are shown for both modified and unmodified lignin samples, and it can be
seen that in the spectra of modified lignin samples, a new set of peaks has appeared in the
950 – 900 cm-1 region. Several researchers report that an oxirane peak appears approximately
840 cm-1 - 915 cm−1 in the FTIR spectra of epichlorohydrin modified lignins and can be
attributed to symmetrical and asymmetrical stretching modes [41- 43]. Others (Malutan et al.
2001) report a slightly higher value of 956 cm-1 from vibration of the epoxy group [44]. In
epichlorohydrin modified coir pith lignin extracted by the alkaline and PEG protocols, new

13
ACCEPTED MANUSCRIPT

peaks have appeared at 963, 927 849, 745 cm-1 and at 939, 882, 833 and 747 cm-1,
respectively. The appearance of these new peaks indicates that an epoxidation reaction has
proceeded. The sharpness of the peaks is an indication that the extent of the epoxidation of
lignin is relatively high. Due to the different extent of hydrogen bonding, the characteristic
epoxy peak’s position has been slightly changed. Moreover, in alkaline lignin, the phenyl
hydroxyl group peak at 3404 cm-1 has shifted to a lower value of 3312 cm-1, and in PEG
lignin, this -OH peak has shifted from 3424 cm-1 to 3358 cm-1. The peak at 1049 cm-1

PT
corresponds to C-O stretching and C-O deformation due to the C–OH group has shifted to
1042 cm-1 in alkaline coir pith lignin. In conclusion, these new structural features indicate that
modified lignins are richer in functional groups than their corresponding raw counterparts.

RI
SC
3.4 Thermal analysis
Thermal data of materials can indicate their complexity; therefore, all lignin samples obtained
NU
by the different extraction protocols were subjected to thermal analysis. Typical analyses of
%wt loss with variation in temperature (TGA) data for selected lignins are shown in Fig. 4.
Derivative temperature difference (°C/°C) with variation in temperature and variation in heat
MA

flow (mW) with variation in temperature (DSC) plots are also included (Fig. S1 and Fig. S2,
supplementary material, respectively). Simultaneous measurements of multiple thermal
D

parameters can strongly improve productivity and simplify the interpretation of results. The
TE

wealth of information produced by these measurements enables differentiation between


thermal events that have no accompanying weight change, such as melting and phase
changes, and those which do involve weight change, such decomposition. It can be seen that
P

the TGA patterns for lignins extracted from the different extraction protocols are quite
CE

different (Fig. 4). The differences can be attributed to differences in the functional groups of
lignins as generated by the various protocols, and due to numerous interactions between
AC

molecules and analogous chemical structures [44]. The onset temperature was taken at 5 %wt
loss, where the polymer material starts oxidative degradation (T5 %), and the temperature at
50 %wt loss (T50 %) and the residual mass at 400 °C were obtained from the TGA data
analysis and are presented in Table 2.

Lignin degradation occurs over a wide range of temperatures due to its complexity, with
many interconnecting phenolic rings with various branches, and it can be divided into
multiple stages. Several authors have suggested that the degradation of lignin occurs in three
stages [46 - 48], while others have suggested four stages [49 - 52]. In the three-stage

14
ACCEPTED MANUSCRIPT

approach, the first clearly visible peaks appear between 30 - 120 °C due to the loss of inner-
bound adsorbed moisture and the decomposition and high-volatile compounds. The second
stage occurs due to the elimination of volatile gases and low-molecular-weight compounds by
the degradation of carbohydrates, α and β aryl-alkyl-ether linked components of lignins, in
the 180 - 350 °C temperature range. Third stage occurs above 350 °C, over a wide range, due
to the elimination of degraded volatile poly-oligomer products derived from lignin [47]. In a
four-stage degradation approach, the sequences of events that occur are similar to the three-

PT
stage degradation process. The first stage occurs in the temperature range between 30 and
130 °C. The weight loss in this stage are primarily attributed to ether linkages, side chain
oxidation (i.e., carbonylation/carboxylation of aliphatic hydroxyl group, side chain

RI
dehydrogenation), and partial C–C (5-5) bond cleavages within lignin biopolymers. The

SC
second stage occurs at 130 - 200 °C, the third stage at 200 - 500 °C, and the fourth stage
above 500 °C. At temperatures higher than 500 °C, many of the high-molecular-weight
NU
compounds with higher cracking energy are decomposed and eliminated. Decomposition
occurs due to deformation reactions involving low bond energy β-O-4 ether bridging
structures in lignin, as well [53]. Gaseous products include carbon monoxide, carbon dioxide
MA

and methane, as analysed by GC-MS [54, 55].


D

Microwave heating promotes the cleavage of the Caryl–Cα bond [23]. Both thermally and
TE

microwave assisted PEG incorporated lignins have shown high weight loss rates (i.e., high
decomposition rates), as shown in Fig. 4. It has been reported that C–C bond cleavage is
difficult to achieve under thermal heating conditions but occurs comparatively easily under
P

microwave conditions and is thereby more mono-phenolic (i.e., low molecular weight lignin
CE

compounds are formed) [53]. Therefore, PEG incorporation into lignin seems likely to lower
its thermal stability, irrespective to the mode of energy supply in the extraction process (Fig.
AC

4). Furthermore, it can be seen that the microwave-assisted PEG-incorporated lignin TGA
curve appears above the thermal energy assisted curve. Thus, it is possible to conclude that
microwave heating with PEG incorporation into lignin extraction caused the lowest thermal
stability. This fact demonstrates that the energy supply mode employed for lignin extraction
directly influences its thermal stability.

The lowest onset degradation temperature of 64 ±1 °C belongs to organosolv lignin, followed


by alkaline lignin, which was 70 ±1 °C. Surprisingly, both PEG lignins have comparatively
high onset degradation temperatures. In PEG-incorporated thermal-assisted lignin, the onset

15
ACCEPTED MANUSCRIPT

degradation temperature was 167 ±2 °C, and this value was lowered by 5 ±2 °C when the
mode of energy supply was changed to microwave. TGA data analysis reveals that alkaline
lignin was the most hygroscopic followed by organosolv lignin, in which 9.33 ±1.12 and 6.42
±0.41 %wt losses, respectively, have taken place in the temperature range of 30 - 120 °C.
Both PEG lignins have an average moisture content of approximately 5 %wt. In the 120 - 200
°C range, the PEG lignins degraded very slowly, indicating ~ 1 %wt loss. Over the same
temperature range (120 - 200 °C), organosolv lignin lost 2.34 ±0.24 %wt and alkaline lignin

PT
lost 6.17 ±1.21 %wt. Overall, the highest %wt loss occurred in the 300 - 400 °C temperature
range, where 9.28 ±0.82, 21.02 ± 1.14, 32.59± 2.32 and 29.79 ± 2.36 %wt losses occurred in
alkaline, organosolv, PEG thermal, and PEG microwave, respectively.

RI
SC
In the temperature range above 100 °C, scattered and fluctuated curves appeared for all
derivative temperature differences (°C/°C) for lignin samples with variation in temperature
NU
curves (Fig. S1 (a) - (e), supplementary materials), due to the slight differences in chemical
structure between lignin biomolecules, which may have been caused by differences in the
chemical extraction protocols. When the ash contents of lignin samples are compared, as in
MA

Table 1, PEG-incorporated lignins have lower values, and microwave-assisted PEG lignin
has the lowest recorded ash content at 2.1 ±0.3 %wt. Generally, a lower residue weight is an
D

indication of a sample’s low thermal stability and an indication of the presence of less
TE

inorganic matter as well.

The alkaline extracted lignin shows a nearly linear rate of %wt loss (Fig. 4). The mean ash
P

content of alkaline lignin is 10.4 ±1.4 %wt at 575 °C, which is the highest of all samples. In
CE

alkaline lignin treated at 800 °C, ~ 60 %wt still remained in an un-volatilized state, signifying
the greatest thermal stability. High non-volatile matter content may be due to the presence of
AC

sodium salts and other inorganic matter [33]. Organosolv lignin gave the second highest
mean ash content of 6.1 ±0.6 %wt, and at 800 °C approximately 45 %wt still remained. The
thermal properties of lignin are important for its use as a filler in other natural and synthetic
materials and to secure a lengthy service life [45, 49].

3.5 DSC analysis of lignin and modified lignin


The DSC curves of selected coir pith lignins extracted by various chemical protocols show
endothermic transitions (Fig. S2, supplementary materials). The heat flow plots of PEG
lignin, extracted either by thermal or microwave energy, lie in a very narrow range and are

16
ACCEPTED MANUSCRIPT

almost overlapping. Lignins possess inherent complex molecular structures, due to different
aromatic rings with various branches that behave as thermosetting polymers. A unique and
characteristic feature of lignin is termed ‘sintering’, in which irregular bonding occurs
between its own functional groups that have no free volume (the space in a solid or liquid
sample which is not occupied by polymer molecules). Therefore, lignin shows no specific
glass transition temperature or melting temperature. The different heat flow patterns observed
may be due to the differences in functional groups caused by different extraction protocols

PT
[54].

RI
4. Conclusions
The effects of different reagents and varying conditions on lignin extraction yields (alkaline,

SC
organosolv and PEG) were studied using coconut coir pith as the biomass source.
Corresponding %wt yields and ash contents were compared with the variation in structural
NU
integrity of lignin based on the mode of energy supply (thermal vs. microwave). Lignin
extraction from rice straw was also attempted. Alkaline lignin extraction with 7.5 % (w/v)
NaOH at 90 °C for 90 min. gave the highest yield of 26 %wt, whereas PEG had the lowest
MA

yield of 2 %wt and organosolv extraction (85 % (v/v) formic acid, 85 % (v/v) acetic acid)
was ~ 6 %wt. The use of microwave radiation drastically reduced the reaction time of the
D

process with the elimination of the need for typical glassware and processing, and it also
TE

significantly increased lignin extraction yields. It was observed that the source of biomass,
the extraction protocols, and the mode of energy supply greatly affected the structural
integrity of the isolated lignins. PEG microwave extracted lignin had the highest degradation
P

rate, as shown in the TGA. Alkaline extracted coir pith lignin was the most thermally stable
CE

lignin, while microwave energy assisted PEG lignin was the least thermally stable. A semi-
quantitative method was developed for the assessment of functional groups based on FTIR
AC

peak height ratio measurement, which determined that in coir pith, both the organosolv and
PEG extraction protocols had a relatively higher extent (~11 %) of carbonyl groups.
Furthermore, additional research on the physicochemical, spectral, and thermal properties of
modified and raw lignin is needed, as thermal and mechanical properties are highly important
factors for the advanced usage of lignin in polymeric applications.

17
ACCEPTED MANUSCRIPT

5 Acknowledgements
The authors are grateful to Prof. K. M. Nalin de Silva and Ms. Niranjala Fernando of Sri
Lanka Institute of Nanotechnology (SLINTEC) for TGA spectra, Prof. L. Karunanayake and
Dr. A. Cooray, Department of Chemistry, University of Sri Jayewardenepura for FTIR
spectra, and Pulasthi Panamgama, University of Moratuwa for graphics and illustrations.

Funding Source

PT
This research was supported by the University of Ruhuna, Science Faculty Research Grant
No. RU/SF/RP/2015/07.

RI
SC
NU
MA
D
P TE
CE
AC

18
ACCEPTED MANUSCRIPT

References
[1]. http://www.srilankabusiness.com/coconut/ (Assessed 07-09-2017).
[2]. http://www.srilankabusiness.com/coconut/why-sri-lankan-coconut-and-coconut-based-
products.html) (Assessed 07-09-2017).
[3]. http://www.srilankabusiness.com/coconut/sri-lankan-coconut-and-coconut-based-
products.html (Assessed 07-09-2017).
[4]. A. L. P. Vidhana, L. L. W. Somasiri, Use of coir dust on the productivity of coconut on

PT
sandy soils, Cocos, 12 (1997) 54–71.
[5]. M. R. Evans, S. Knoduru, R. H. Stamps, Source variation in physical and chemical
properties of coconut coir dust, HortScience 31 (1996) 965-967.

RI
[6]. A. W. Mererow, Growth of two sub-tropical ornamental plants using coir (coconut

SC
mesocarp pith) as a peat substitute, HortScience 29 (1994) 1484–1486.
[7]. E. T Joseph, V. C. Sarma, Coconut fibre morphology and chemical composition, J. Food
NU
Sci., 67(1) (1997) 420-424.
M. Abad, P. Nougera, R. Punchades, A. Maqueira, V. Nouguera, Physicochemical and
chemical properties of some coconut coir dust for use as a peat substitute. Bioresour.
MA

Technol. 83(2) (2002) 241-245.


[8]. A. Garcia, Toledano A. Serrano L. Egues, I. M Gonzalez, F. Martin J. Labidi,
D

Characterization of lignins obtained by selective precipitation, Sep. Purif. Technol, 68


TE

(2009) 193–198.
[9]. Y. P. Timilsena, I.G. Audu, S. K. Rakshit, N. Brosse, Impact of the lignin structure of
three lignocellulosic feedstocks on their organosolv delignification: Effect of carbonium
P

ion scavengers, Biomass and Bioenergy, 52 (2013) 151-158.


CE

[10]. Q. Yin, W. Yang, C. Sun, M. Di, Preparation and properties of lignin-epoxy resin
composite, BioResources, 7(4) (2012) 5737-5748.
AC

[11]. S. Laurichesse, L Averous, 2014. Chemical modification of lignins, Prog. Polym. Sci., 39
(2014) 1266-1290.
[12]. L. Hu, H. Pan, Y. Zhou, M. Zhang, Methods to improve lignin’s reactivity as a phenol
substitute and as replacement for other phenolic compounds: A brief review.
BioResources, 6(3) (2011) 3515-3525.
[13]. A.R. Goncalves, P. Benar, Hydroxymethylation and oxidation of organosolv lignins and
utilization of the products, Bioresour. Technol., 79 (2001) 103-111.
[14]. P. Raquel, E. Xabier, L. Jalel, Lignin extraction and purification with ionic Liquids, J
Chem Technol Biotechnol, 88 (2013) 1248–1257.

19
ACCEPTED MANUSCRIPT

[15]. Y. Matsushita, Conversion of technical lignins to functional materials with retained


polymeric properties, J. Wood Sci. 61 (2015) 230-250.
[16]. D. Watkins, M. Nuruddin, M. Hosur, A. T. Narteh, S. Jeelani, Extraction and
characterization of lignin from different biomss resources, J Mater Res Technol. 4(1)
(2015) 26-32.
[17]. M. E. Vallejos, F. E.Fellissia, A. A. S Curvelo, M. D.Zambon, L. Ramos, M. C. Area,
Chemical and physico-chemical characterization of lignins obtained from ethanol-water

PT
fractionation of bagasse, BioResources 2 (2011) 1158-1171.
[18]. R. Bayerbach, D. Meier, Characterization of the water-insoluble fraction from fast
pyrolysis liquids (pyrolytic lignin). Part IV: structure elucidation of oligomeric molecules.

RI
J. Anal. Appl. Pyrol. 85 (2009) 98–107.

SC
[19]. F. M. Rivera, M. Phuong, M. Ye, A. Halasz, Isolation and characterization of herbaceous
lignin for applications in biomaterials. Ind. Crops Prod. 41 (2013) 356–364.
NU
[20]. H. J. G. Jung, D. S. Himmelsbach, Isolation and characterization of wheat straw lignin, J.
Agric. Food Chem. 37 (1989) 81-87.
[21]. F. M. Rivera, G. H. Huang, L. Paquet, S. Deschamps, C. Beaulieu, Microwave-assisted
MA

extraction of lignin from triticale straw: Optimization and microwave effects, Bioresour.
Technol., 104 (2012) 775-782.
D

[22]. O. Xinping, Z. Guodian, H. Xiangzhen, Q. Xueqing, Microwave assisted liquefaction of


TE

wheat straw alkali lignin for the production of monophenolic compounds, J Energy
Chem. 24 (2015) 72–76.
[23]. K. B. Kaufmann, C. Philippe, Recent extraction techniques for natural products:
P

microwave-assisted extraction and pressurized solvent extraction, Phytochem. Anal. 13


CE

(2002) 105–113.
[24]. L. Yanming, L. Bingzheng, D. Fangli, W. Yong, P. Lixia, C. Dong, Microwave-assisted
AC

hydrothermal liquefaction of lignin for the preparation of phenolic formaldehyde


adhesive, J. Appl. Polym. Sci., 134 (2017) 44510.
[25]. L. Gaiyun, H. Chungyun, Q. Tefu, Wood liquefaction with phenol by microwave heating
and FTIR evaluation J. For. Res. 26(4) (2015) 1043–1048.
[26]. M.R. Fanny, P. Marielle, Y. Mengwei, H. Annamaria, H. Jalal, Isolation and
characterization of herbaceous lignins for applications in biomaterials, Ind. Crops Prod.,
41 (2013) 356 – 364.

20
ACCEPTED MANUSCRIPT

[27]. A. Tejado, C. Pen, J. Labidi, J.M. Echeverria, I. Mondragon, Physico-chemical


characterization of lignins from different sources for use in phenol–formaldehyde resin
synthesis, Bioresour. Technol., 98 (2007) 1655–1663.
[28]. F. Xu, J. X.Sun, R. C. Sun, P. Fowler, M. S. Baird, Comparative study of organosolv
lignins from wheat straw. Ind. Crops Prod. 23 (2006) 180–193.
[29]. M. F. Li, S.N. Sun, F. Xu, R.C. Sun, Formic acid based organosolv pulping of bamboo
(Phyllostachys acuta): comparative characterization of the dissolved lignins with milled

PT
wood lignin, Chem. Eng. J., vol. 179 (2012) 80–89.
[30]. Z. Xuebing, L. Dehua Chemical and thermal characteristics of lignins isolated from Siam
weed stem by acetic acid and formic acid delignification, Ind. Crops Prod., 32 (2010)

RI
284–291.

SC
[31]. M. G. Alriols, A. Tejado, M. Blanco, I. Mondragon, J. Labidi, Agricultural palm oil tree
residues as raw material for cellulose, lignin and hemicelluloses production by ethylene
NU
glycol pulping process, Chem. Eng. J., 148 (2009) 106–114.
[32]. J.B. Sluiter, R.O. Ruiz, C.J. Scarlata, A.D. Sluiter, D.W. Templeton, Compositional
Analysis of Lignocellulosic Feedstocks. 1. Review and Description of Methods. J. Agric.
MA

Food Chem. 58 (2010) 9043–9053.


[33]. O. Faix, Classification of lignins from different botanical origins by FTIR
D

spectroscopy. Holzforschung, 45 (1991) 21-28.


TE

[34]. J. G. Buta, F. Zadrazil, G. C. Galletti, FTIR Determination of lignin degradation in wheat


straw by white rot fungus Stropharia rugosoannulata with different oxygen
concentrations. J. Agric. Food Chem. 37 (1989) 1382-1384.
P

[35]. C. G. Boeriu, D. Bravo, R. J. A., Gosselink, J. E. G. v. Dam, Characterizations of


CE

structure-dependent functional properties of lignin with infrared spectroscopy. Ind. Crops


Prod. 20 (2004) 205–218.
AC

[36]. O. Faix, O. Beinhoff, FTIR spectra of milled wood lignins and lignin polymer models
(DHP's) with enhanced resolution obtained by deconvolution, J wood chem. Tech., 8, 4
(1988) 505-522.
[37]. A.J. Michelle, A.J. Watson, H.G. Higgins, An Infrared spectroscopic study of
delignification Eucalyptus regnans. Tappi 48(9) (1965) 520–532.
[38]. S. Kubo, J.F. Kadla, Hydrogen bonding in lignin: A Fourier transform infrared model
compound study. Biomacromolecules 6 (2005) 2815–2821.

21
ACCEPTED MANUSCRIPT

[39]. J. D. Robles, R. Sánchez, E. Espinosa, D. Savy, P. Mazzei, A. Piccolo, A.Rodríguez,


Isolation and characterization of gramineae and fabaceae soda lignins, Int. J. Mol. Sci. 18
(2017) 327 doi.org/10.3390/ijms18020327.
[40]. F. Fatemeh, Y. Zhongshun, M. Andersonb C. Xu, Synthesis of lignin-based epoxy resins:
optimization of reaction parameters using response surface methodology RSC Adv., 4
(2014) 31745–31753.
[41]. F. Fatemeh, Y. Zhongshun, A. Mark, C. Xu, Sustainable lignin-based epoxy resins cured

PT
with aromatic and aliphatic amine curing agents: Curing kinetics and thermal properties
Thermochimica Acta 618 (2015) 48–55.
[42]. R.G.C. Arridge, J.H. Speake, Mechanical relaxation studies of the cure of epoxy resins: 1.

RI
Measurement of cure, Polymer, 13 (1972) 443-449.

SC
[43]. T. Malutan, R. Nicu, V.I. Pops, Lignin modification by epoxidation, Bioresources, 3 (4)
(2008) 1371-1376.
NU
[44]. Q. Liu, S.R. Wang, Y. Zheng, Z.Y. Luo, K.F. Cen, Mechanism study of wood lignin
pyrolysis by using TG–FTIR analysis. J. Anal. Appl. Pyrol. 82 (2008) 170–177.
[45]. A. Moubarik, N. Grimi, N. Boussetta, A. Pizzi, Isolation and characterization of lignin
MA

from Moroccan sugar cane bagasse: Production of lignin–phenol-formaldehyde wood


adhesive, Ind. Crops Prod. 45 (2013) 296– 302.
D

[46]. T. Faravelli, A. Frassoldati, G. Migliavacca E. Ranzi, Detailed kinetic modeling of the


TE

thermal degradation of lignins. Biomass Bioenergy 34 (2010) 290–301.


[47]. A. Toledano, L. Serrano, A. Garcia, I. Mondragon, J. Labidi, Comparative study of lignin
fractionation by ultrafiltration and selective precipitation, Chem. Eng. J. 157 (2010) 93–
P

99.
CE

[48]. W.O.S. Doherty, P. Mousavioun, C.M. Fellows, Value-adding to cellulosic ethanol lignin
polymers. Ind. Crop. Prod. 33 (2011) 259–276.
AC

[49]. O. Gordobil, R. Moriana, L. Zhang, J. Labidi, O. Sevastyanova, Assessment of technical


lignins for uses in biofuels and biomaterials: Structure-related properties, proximate
analysis and chemical modification, Ind. Crop. Prod. 83 (2016) 155–165.
[50]. R. Sun, Q. Lu, X.F.Sun, Physico-chemical and thermal characterization of lignins from
Caligonum monogoliacum and Tamarix spp. Polymer Degr. Stab. 72 (2), (2001) 229–
238.
[51]. M. Brebu, C. Vasile, Thermal degradation of lignin – A Review. Cell. Chem. Technol.
(2009) 44, 353.

22
ACCEPTED MANUSCRIPT

[52]. J. Y. Kim, H. Hwang, S. Oh, Y. S. Kim, U. J. Kim, J. W. Choi, Investigation of structural


modification and thermal characteristics of lignin after heat treatment. Int. J. Biol.
Macromol. 66 (2014) 57. DOI: 10.1016/j.ijbiomac.2014.02.013
[53]. S.N. Monteiro, V. Calado, F.M. Margem, R.J.S. Rodriguez, Thermogravimetric stability
behavior of less common lignocellulosic fibers – a review, J. mater. res. technol. 1 (2012)
189–199.
[54]. S. Kim, J. Park, J. Lee, H. Roh, D. Jeong, S. Choi, S. Oh, Potential of a bio-disintegrable

PT
polymer blend using alkyl-chain-modified lignin, Fibe. Poly., 16 (4) (2015) 744-751.
DOI: 10.1007/s12221-015-0744-z.

RI
SC
NU
MA
D
P TE
CE
AC

23
ACCEPTED MANUSCRIPT

Table 1. Mean %wt lignin yields and ash contents obtained for different extraction
protocols

Energy Biomass Extracting Mean Lignin Mean Ash


Source Solvents Yield (%) Content (%)
Alkaline 26.4± 1.5 10.4± 1.4
Thermal Coir Pith Organosolv 3.4 ± 0.2 6.1 ± 0.6

PT
PEG 6.7 ± 0.6 2.2 ± 0.3
Rice Straw Alkaline 3.1 ± 0.3 11.1± 1.2

RI
Alkaline 32.4 ± 1.6 10.4 ± 0.8
Microwave Coir Pith

SC
Organosolv 4.1 ± 0.4 6.4 ± 0.5
PEG NU 7.2 ± 0.8 2.1 ± 0.3
MA
D
P TE
CE
AC

24
ACCEPTED MANUSCRIPT

Table 2. Thermal decomposition characteristics of lignin obtained by different


extraction protocols from TGA data
Energy Source

120–200 °C (%)

300–400 °C (%)
27–120 °C (%)

Residual yield
T5 oC (%)

at 400 °C wt.
T50 oC (%)
200 -300 oC
Mass loss at

Mass loss at

Mass loss at

Mass loss at
Lignin

(%)
(%)

PT
Alkaline 9.33±1.22 6.17±1.21 11.2±2.11 9.28±0.82 70±1 559±4 47±1
Thermal

Organosolv 6.42±0.41 2.34±0.24 16.15±0.23 21.02±1.14 64±1 433±3 46±1

RI
PEG 4.67±0.22 1.31±0.12 12.08±2.4 32.59±2.32 167±2 396±2 51±2

SC
Microwa

PEG 4.63±0.34 1.14±0.08 12.29±2.62 29.76±2.36 162±2 - 48±2


ve

NU
MA
D
P TE
CE
AC

25
ACCEPTED MANUSCRIPT

Table 3. Results of semi-quantitative relative peak height ratio data for selected
functional groups of lignins obtained by different extraction protocols

Extraction protocol Alkaline Alkaline PEG Organosolv Modification with


Epichlorohydrin
Lignin variety Straw Coir Pith
PEG Alkaline

PT
I Carbonyl/IC=C 0.97 0.93 1.05 1.05 1.01 0.98
I Methoxyl/IC=C 0.97 0.7 0.99 1.03 0.98 0.89

RI
SC
N.B. Calculation is based on the C=C phenyl skeletal vibration peak at 1510 cm-1 as the
reference peak in FTIR spectroscopy.
NU
MA
D
P TE
CE
AC

26
ACCEPTED MANUSCRIPT

Figure Captions

Fig. 1 Coir pith blocks (a) top view and (b) side view

Fig. 2 (a) FTIR spectra of selected thermal energy assisted extraction lignin samples
wavenumber from 4000 to 400 cm-1 range

PT
Fig. 2 (b) FTIR spectra of selected thermal energy assisted extraction lignin samples
magnified region of 2000 to 400 cm-1

RI
Fig. 3 (a) FTIR spectra of alkaline coir pith lignin and epichlorohydrin modified alkaline coir

SC
pith lignin obtained by thermal energy assisted extraction.
NU
Fig. 3 (b) FTIR spectra of PEG coir pith lignin and epichlorohydrin modified PEG coir pith
lignin by thermal energy assisted extraction.
MA

Fig. 4 Thermogravimetric curves (TGA) of selected coir pith lignin obtained from different
chemical extraction protocols and energy supply modes under nitrogen atmosphere at 10
D

°C/min
TE

Fig. S1 (a) - (e) Derivative temperature difference (°C/°C) of a lignin sample with variation in
temperature of selected coir pith lignin obtained from different chemical extraction protocols
P

and energy supply modes under nitrogen atmosphere at 10 °C/min


CE

Fig. S2 DSC curves of selected coir pith lignin obtained from different chemical extraction
AC

protocols.

27
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 1 Coir pith blocks (a) top view and (b) side view.

28
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 2 (a) FTIR spectra of selected thermal energy assisted extraction lignin samples
wavenumber from 4000 to 400 cm-1 range

29
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 2 (b) FTIR spectra of selected thermal energy assisted extraction lignin samples
magnified region of 2000 to 400 cm-1

30
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA

Fig. 3(a) FTIR spectra comparison of alkaline coir pith lignin and epichlorohydrin
modified alkaline coir pith lignin obtained by thermal energy assisted extraction.
D
P TE
CE
AC

31
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
TE

Fig. 3(b) FTIR spectra comparison of PEG coir pith lignin and epichlorohydrin
modified PEG coir pith lignin by thermal energy assisted extraction.
P
CE
AC

32
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
Fig. 4 Thermogravimetric curves (TGA) of selected coir pith lignin obtained from
MA

different chemical extraction protocols and energy supply modes under nitrogen
atmosphere at 10 °C/min.
D
P TE
CE
AC

33

You might also like