You are on page 1of 10

International Journal of Adhesion & Adhesives 48 (2014) 92101

Contents lists available at ScienceDirect

International Journal of Adhesion & Adhesives


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

Lignin in straw and its applications as an adhesive


Seyed Hamidreza Ghaffar, Mizi Fan n
Nanocellulose and Biocomposites Research Centre, School of Engineering and Design, Brunel University, Uxbridge, Middlesex UB8 3PH, United Kingdom

art ic l e i nf o

a b s t r a c t

Article history:
Accepted 30 May 2013
Available online 27 September 2013

The relevant information about the lignin in straw and its applications in the industry is scattered and
scarce compared to the wood lignin. This review is focused on the chemical structural and composition of
lignin in the straw, and its modication and uses as an adhesive. The review has showed that (1) lignin as
a by-product in the pulping process and as an abundant natural and renewable product has been used
and there is a great potential for many applications across various industrial sectors as a replacement for
increasingly scarce and expensive petroleum based materials, including traditional products, e.g. resins,
and composites, and emerging materials, e.g. biofuel and commodity chemicals. (2) The type of lignin
differs not only from one to another species but also depending on the isolation protocol. However, the
lack of optimising or processing technologies is signicant when it comes to using technical lignin. The
review has also shown a great encouragement in studying the lignin within the straw and other
herbaceous crops, and the creation of the functionalities of lignin as it does with cellulose and
hemicellulose could lead to radical development of lignin as bio-matrix for green composites and
biomass as biofuel or other high value added applications.
& 2013 Elsevier Ltd. All rights reserved.

Keywords:
Lignin
Straw
Adhesive
Lignocellulosic
Lignin-based products

1. Introduction
Straw is generally constituted of three groups of organic mixtures:
cellulose, hemicellulose and lignin. These compounds account for
more than 80% of the dry matter of the three most common British
cereal straws: namely barley, oats and wheat [1]. The chemical
content of straw varies according to the stage of maturity, soil type
and fertiliser treatment Table 1 [2].
Straw also contains various other organic compounds including
protein, small quantities of waxes which protect the epidermis of
the straw, sugars, salts and insoluble ash including silica. The lignin
content for those main straw species ranges from 12 to 21% although
it varies from one to another species Table 1. The rst discovery of
lignin was recorded from 1838 by Anselme Payen cited by McCarthy
and Islam [3]. However, the word lignin was put forward in 1865 by
Schulze to describe the dissolved part of wood when treated with
nitric acid [4,5]. The research on lignin has proceeded at a fast pace
since 1960s when powerful modern analytical tools of biochemistry
and organic chemistry were applied and therefore interesting
information was gathered. Research on lignin has also attracted
much attention because of the dominant pulping industries [3] and
recently a large number of studies on the lignin focused on nding a
higher value application [4].

n
Correspondence to: Department of Civil Engineering, Brunel University,
UB8 3PH United Kingdom. Tel.: 44 1895 266466.
E-mail address: mizi.fan@brunel.ac.uk (M. Fan).

0143-7496/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijadhadh.2013.09.001

Lignin is an amorphous natural polymeric material that is


based on phenylpropane derivate and one of the most abundant
materials and renewable resources on earth. It is well accepted
that lignin is a phenolic polymer derived primarily from three
hydroxycinnamyl alcohols or monolignols by free radical generation followed by chemical coupling process. The hydrocycinnamyl
are p-coumaryl alcohol (MH), coniferyl alcohol (MG) and sinapyl
alcohol (MS) [6] (Fig. 1).
Many attempts have been made to dene lignin based on the
constitution, structural features and mechanism of formation.
However, the denition of lignin has never been as clear as that
of other natural polymers such as cellulose and protein, the reason
being the extremely complicated isolation, compositional analysis
and structural characterisation [6]. Lignin is never dened as a
compound, is a class of phenolic natural polymers with broad
compositions and a variety of linkages between units. The problem
of lack of precise denition for lignin is also associated with its
nature; no regularly repeating multiunit structures have been
found, and composition and structure of lignin vary depending
on their origins [6]. In a recent review of lignin, a denition given
by Brunow was cited as the most concise and comprehensive one
to date [7]: lignins are biopolymers consisting of phenylpropanoid
units with an oxygen atom at the p-position (as HO or OC) and
with none, one or two methoxyl groups in the o-position to this
oxygen atom.
There are many review articles and book chapters available that
focus on structure, composition and biosynthesis of lignin, which
are mostly wood lignin [4,8,9]. Reviews on lignin in herbaceous

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

Table 1
Chemical composition of agricultural residues (% dry matter) based on [2].
Type

Lignin Cellulose Hemicellulose Water- Wax Ash


soluble

Others

Wheat straw
Rice straw
Rye straw
Barley straw
Oat straw
Rape straw
Maize stems
Corn cobs
Esparto
Bagasse
Rye grass
Oil palm bre
Abaca bre

14.1
12.3
17.6
14.6
16.8
21.3
15.0
14.6
17.8
19.4
8.2
18.7
12.4

2.4
0.3
2.6
8.3
3.1
0
5.1
0.1
1.7
2
4.6
0.1
0

38.6
36.5
37.9
34.8
38.5
37.6
38.5
43.2
35.8
39.2
37.6
40.2
60.4

32.6
27.7
32.8
27.9
31.7
31.4
28.0
31.8
28.7
28.7
32.2
32.1
20.8

4.7
6.1
4.1
6.8
4.6
5.6
4.2
6.1
4.0
8.5
5.0
3.7

1.7
3.8
2.0
1.9
2.2
3.8
3.6
3.9
3.4
1.6
4.4
0.5
0.8

5.9
13.3
3.0
5.7
3.1
6.0
4.2
2.2
6.5
5.1
4.5
3.4
2.5

93

of binder while ensuring product quality is vital for the industry. The
transition from petroleum based chemical processes to bioprocesses
based on renewable raw materials provides considerable environmental gains and such the maintenance of economic growth.
The oil crisis during the 1970s turned attention towards the
utilisation of renewable resources and towards lignocellulosic
materials in specic, hence biotechnical utilisation of lignocellulosic
wastes from agriculture and forestry gained priority. This was a
logical step to take since one of nature's most important biological
reactions is the conversion of wood and other lignocellulosic
materials to carbon dioxide, water and humic substances [10]. Since
lignocellulosic plants constitute the raw material for the forest
industries, there should be plenty of opportunities to use biotechnology to improve both the production and the use of these
resources. Biotechnical utilisation and conversion of lignocellulosic
materials mean the production of inexpensive products on a large
scale. Biotechnological utilisation of lignocellulosic materials is
therefore a very difcult task and the commercial utilisation of this
technology has only recently gained momentum. Lignin in straw is
receiving increasing attentions, the reason being its annual renewability and herbaceous plants have the largest annual biomass stock
(1549 million tons/year worldwide) [11]. The main difculty for the
utilisation of herbaceous crops biomass is the complete separation
of lignincarbohydrates complexes (LCC) which shield cellulose
from enzymatic hydrolysis and fermentation. The enzymatic digestibility of the biomass for production of bio products and biofuels
depends mainly on its lignin content [12,13]. Consequently, the
extraction of lignin from straw is important and the utilisation of
extracted lignin could lead to the industrial production of treasured
industrial products such as vanillin, ferulic acid, vinyl guaiacol and
optically active lignans, the dimers of monolignols.

3. Types and variations of straw lignin

Fig. 1. Three primary lignin monomers: (a) monolignols p-coumaryl alcohol,


(b) coniferyl alcohol and (c) sinapyl alcohols [7].

plants are limited. It is evident that the interest in herbaceous


crops and their lignin is increasing and will increase further due to
the importance of bioethanol production from annually renewable
biomass. Hence this paper reviews the structural characterisation
of lignin in straw and the process for purication and isolation of
lignin. The focus of the review is centered on the applications of
lignin in the engineering industry, e.g. for the development of
construction materials, such as the use of lignin as adhesives for
strawboard, particleboard, breboard, oriented strand board and
wood bre insulation board.

2. Why utilising lignin in engineering?


Renewable resources in engineering industry are very essential,
and the creative design and wise application of innovative technology
for the optimisation of such resources is one of the new research
topics. Using a system-based approach to design technology for the
sustainable development, management and conservation of our
natural resources is what researchers are keen to develop. Using
renewable resources in many aspects means a reduction of energy
consumed synthesis and harmful substances from end product. For
example manufacturers of medium density breboard (MDF), particleboard (PB), plywood and oriented strand board (OSB) are under
pressure to reduce production costs and harmful formaldehyde
emissions from the petroleum derived adhesives, and to improve
product recyclability. Innovative approaches to minimise the amount

Lignin is found as a cell wall component in all vascular plants [4].


The lignin content in plant stems varies between 15 and 40%. Lignin
acts as water sealant in the stems and plays an important role in
controlling water transport through the cell wall. It also protects
plants against biological attack by hampering enzyme penetration
and acts as permanent glue, bonding cells together in the plant stems
and thus giving the stems their well-known rigidity and impact
resistance [14]. Most industrial lignin is obtained as a waste product
during the paper pulping process, but it can be found in all plants.
While cellulose is used for paper production and natural oils are
mainly used in the food industry, the industrial applications of lignin
are rather limited, despite its widespread availability. In 1998 about
1% of all lignin generated in paper production worldwide was isolated
and sold [3]. The remaining 99% was either burned in an energy
recovery step for the pulping process or disposed of in waste streams.
Extrapolating this 1% lignin leads to a worldwide production of more
than 10 million tons per year of available renewable raw material [15].
A number of studies on lignin focused on nding a higher value
application [1618].
The lignin within the straw is known to be structurally different
from hardwood and softwood lignin; one of the signicant
differences is the monomeric composition. Straw lignin comprises
all three H, G and S subunits. On the other hand, wood lignin
contains mainly G and S subunits (except for compression wood
which contains H and G units) [17]. The physicochemical properties of straw lignin are known to possess characteristic alkali
solubility [19]. Alkali treatments have been used to increase the
digestibility of various straws and to manufacture paper [20,21].
The solubility of straw lignin in alkali has been attributed mainly
to the presence of signicant amounts of p-hydroxyphenyl
(H) residues, which are bound to lignin as p-coumarate units

94

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

[19,22]. Ferulic and p-coumaric acids in grasses are known to be


implicated in the cross-linking of cell wall carbohydrates with
lignin [23,24]. Obviously such cross-linking can have a dramatic
inuence on the mechanical properties and biodegradability of
straws. The lignied plant cell walls are developed by successive
deposition of cellulose, hemicellulose and lignin to form a composite in which these components are physically and chemically
connected to each other [6]. For this reason lignin is always linked
with other components especially with carbohydrates (hemicellulose), LCC was rst used by Bjrkman to describe a preparation of
hemicellulose accompanied by lignin [25]. LCC and its related
topics have been reviewed [12,26,27]. The existence of covalent
bonds between lignin and carbohydrates are reported and four
types of linkages between lignin and carbohydrates are projected
as: (1) glycosidic linkages; (2) ester linkages; (3) benzyl ether
linkages and (4) hemiacetal or acetal linkages [6].

all three monolignol units in signicant amounts. The analysis of


degradation products following cleavage of ether linkages by
thioacidolysis indicates the respective proportions of H, G and S
units in lignins are 15, 45 and 40% respectively. The p-hydroxyphenyl content in rice straw is signicantly higher than that of
corn and wheat straw [28,29]. Azuma and Koshimjima [32]
describe the methods of isolation, purication and fractionation
for lignincarbohydrate complexes from 1 kg of rice straw in their
work. The LCC is removed from extractive free and de-pectinated
plant meal with 10 L of 80% aqueous dioxane for 48 h. The analysis
of LCC from rice straw has shown that it contains 63.9% carbohydrates (xylose, 80.1%; arabinose, 13%; glucose, 4.3%; galactose, 2.3;
mannose, 0.4%), 2.8% of uronic acid, 27.7% Klason lignin, 5.6% acidsoluble lignin, 4.2% acetyl group, 4% trans-p-coumaric acid (4%)
and 0.8% trans-ferulic acid [17].
3.3. Lignin in corn stover

3.1. Lignin in wheat straw


Asia is the largest region producing global wheat straw with
43% and then Europe with 32% [11]. Due to renewability and low
lignin content, wheat straw is an attractive low cost feedstock for
production of fuel alcohol. When compared to corn stover, wheat
straw comprises lower amounts of lignin and higher cellulose and
hemicellulose [17]. The analysis of degradation products following
cleavage of ether linkages by thioacidolysis indicates the respective proportions of H, G and S units in wheat sraw lignins are 5, 49
and 46% for wheat straw [28,29]. A specic feature of wheat straw
cell walls is the presence of a non-core lignin which is simply
soluble in alkali and represents up to 20% of total lignin. The
removal of this particular lignin increases the digestibility of the
material and improves further microbial or enzymatic bioconversion [30]. Tapin et al. [31] compared phenolic compounds found in
oilseed ax straw with those in wheat straw (Fig. 2), total content
of phenolic compounds is higher in wheat straw than ax straw.
Except the vanillin content, ferulic acid and coumaric acid content
for wheat straw are much higher than ax straw, the reason being
related to the lignin/phenolicscarbohydrate complexes which
have phenolic bridges between lignin and carbohydrates. This
shows that there are no such phenolic bridges between lignin and
carbohydrates in ax straw.
3.2. Lignin in rice straw
World production of rice straw is 525 million tons per year;
90% of this rice straw is produced in Asia [11]. Most of the rice
straw is burnt which leads to environmental contamination. Rice
straw is unique relative to other cereal straws as it contains lower
lignin but higher silica (ash) [17]. Rice straw lignin is also called
p-hydroxyphenylguaiacylsyringyl (HGS) lignin and contains

Quantity (mg/g)

1.5
Wheat
1
Flax
0.5

0
Ferulic acid

Coumaric acid

Vanillin

Vanillin acid

Fig. 2. Phenolic compounds in ax and wheat straw [31].

Corn stover is the surface material remaining after the grain is


removed. About 520 million tons per year of dry corn is produced
annually in the world. The main production regions are North
America (42%), Asia (26%), Europe (12%) and South America (9%)
[11]. Lignin content of corn stover is 1521%, depending on the
production location, and this is relatively low compared to other
herbaceous crops [17]. The analysis of degradation products
following cleavage of ether linkages by thioacidolysis indicates
the respective proportions of H, G and S units in lignins are 4, 35
and 61% respectively for mature maize stalks [28,29]. In contrast to
wheat straw, the monomeric composition of corn lignin does not
change signicantly during nitrobenzene oxidation at high temperature and reaction time [33].
The fractionation of corn stover has been extensively studied for
production of biofuel and biochemical due to its lower lignin content
and large amount of annual biomass. The availability of signicant
amount of ester bonds in lignin/phenoliccarbohydrate complexes
of corn stover requires the use of alkali for the fractionation process.
Consequently, the fractionation technologies used for corn stover are
based on different alkalis in various physical processes, because of
the ester bonds between arabionoxylan and hydroxycinnamic acids.

4. Chemical structure and distribution of straw lignin


The studies on lignin could be divided into two different
sections: qualitative and quantitative analyses. The qualitative
analysis of lignin generally focuses on the H/G/S ratio and the
nature of the bonds using destructive methods, such as nitrobenzene oxidation [34,35], which give rise to patterns in empirically
known degradation products. Lignin composition and H/G/S ratios
for different biomass resources are listed in Table 2.
Quantitative analysis is based on gravimetry or UV-absorption [38]
either to estimate lignin as an insoluble residue after strong sulfuric
acid treatment (Klason lignin) or to oxidise away the lignin from a
holocellulose preparation (acidic chlorite lignin or permanganate
lignin) [39]. The spectroscopic techniques, such as infrared (IR) and
13
C nuclear magnetic resonance (13C-NMR) spectroscopy are complementary to the above procedures since they provide information on
the whole structure of the polymer and avoid the possibility of
degradation artefacts [40]. While the chemical structure of wood
lignin is much better known and a number of techniques are available
for the characterisation of wood lignin [41], the studies concerning
structure and properties of straw or grass lignin (Gramineae lignin)
are very scarce [42], neither the structure of straw and grass lignin is
yet fully understood, nor is their exact interrelationships with other
cell-wall components. Lignin's complexity makes the development of
suitable tools for characterising lignin in plant cell walls very difcult,

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

Table 2
H/G/S ratio for types of biomass resources.
Type

Lignin (%)

Reference

Wheat straw
Rice straw
Rye straw
Hemp
Flax

1621
6
18
813
2134

5
15
1
9
4

49
45
43
51
67

46
40
53
40
29

[28]
[17]
[36]
[37]
[17]

following a wide variety of chemical, physical and biological treatments. Analytical methods for lignin characterisation are classied
into two groups: destructive and non-destructive [6]. Analytical
methods for lignin characterisation are essential in understanding
structure and composition of lignins. Lignin's complexity makes the
development of suitable tools for characterising lignin in plant cell
walls very difcult, following a wide variety of chemical, physical and
biological treatments.
A limited studies showed that overall, wheat straw acidic
chlorite lignin accounts for about 14% of the dry matter, which is
one of the three main components of wheat straw after cellulose
and hemicelluloses. Depending on its composition of guaiacyl (G),
syringyl (S) and p-hydroxyphenypropane (H) units, wheat straw
lignin or grass lignin has been justied as GSH-lignins (Gramineae
lignins from grasses), which are known to be different from those
of softwood (G-lignins) or hardwood (GS-lignins) and compression
wood (GH-lignins) lignins [43]. Phenolic acids, mainly p-coumaric
and ferulic acids, have been investigated to cross-links between
lignins and polysaccharides, and that p-coumaric acid (PCA) in
wheat straw is mainly ester linked to the lignin while ferulic acid
(FA) is ether linked to lignin and ester linked to hemicelluloses
[40,44]. Due to the above complex nature of the straw lignin, the
study of its structure has been found to be more difcult. Moreover many plants have lignins containing signicant levels of other
unusual components and it is likely that no plants contains lignins
that are solely derived from the three main precursors, hence
lignin is viewed not as a constitutionally dened compound, but as
a composite of physical and chemically heterogeneous materials,
whose structure may be represented by models such as those
proposed for wheat stray [43] (Fig. 3), which should not be
regarded as depicting the structural formulas for lignin in the
usual sense, but as a medium for illustrating the types and linkage
modes of constituent structural elements and the proportions in
which they are believed to occur in lignin [45].
Lignin concentration is normally higher in the compound
middle lamella (CML) and cell corners (CC) than in S2 secondary
wall [4648]. CC is the highest lignied in cell walls. Nevertheless,
because it occupies a larger portion of the cell wall, most of the
lignin is in the secondary wall. It was reported that the CML region
of the cell typically contains more than 50% lignin concentration
(w/w), while the S2 region contains about 20% [48]. Microautoradiography and ultraviolet (UV)microspectrometric studies
showed that the incorporation of the three monolignols in grass
lignin is spatially and temporally regulated and varies between
primary and secondary cell walls and among tissues [49].
It is always a challenging task to isolate lignin for compositional
and structural analyses, although many protocols have been
developed for this purpose [4,50]. Previously developed methods
for lignin isolation are focused and fully dedicated to wood,
although some of these protocols have been applied to grass
plants with partial accomplishment [50,51]. The most used lignin
preparation for morphological studies are milled wood lignin
(MWL) and cellulolytic enzyme lignin (CEL) [6]. Sun and his coworkers have isolated and characterised over 200 fractions of
polysaccharides and lignin from wheat straw [43,52], and mainly

95

focused on the effects of mechanical pulping [53], mild alkaline


pre-treatment and steam treatments on the polymeric components of the straw.
4.1. Lignin polymer of straw
Lignin is an amorphous polymer consisting of phenylpropane
units, their predecessors are three aromatic alcohols (monolignols)
during the lignication process, these monolignols produce a
complex three-dimensional amorphous lignin polymer via -O-4,
-O-4, -5, -1, 5-5, 4-O-5 and - linkages which lacks the
regular and ordered repeating units found in other polymers such
as cellulose and proteins [12]. This biosynthesis process consists of
mainly radical coupling and creates a unique lignin polymer in
each plant species. Wood lignin mainly contains guaiacyl and
syringyl units, whereas the lignins of herbaceous plants contain
all three units (H, G, S) in signicant amounts with different
ratios [29,33]. Lignin is always associated with carbohydrates
(in particular with hemicelluloses) via covalent bonds at two sites:
-carbon and C-4 in the benzene ring, and this association is called
lignincarbohydrate complexes (LCC). In herbaceous plants,
hydroxycynnamic acids (p-coumaric and ferulic acids) are attached
to lignin and hemicelluloses via ester and ether bonds as bridges
between them forming lignin/phenolicscarbohydrate complexes
[46,54]. Because of this chemical nature of the lignin, it is
practically impossible to extract lignin in pure form.
4.2. Lignin/phenoliccarbohydrate complexes of straw
The lignincarbohydrate complexes from straw are structurally
different from those in woods; they contain ferulic bridges between
lignin and carbohydrates (arabionxylans) through ester-linked
ferulic acids [55]. Therefore, they are often referred to as lignin/
phenolicscarbohydrate complexes (Fig. 4) [43].
The lignincarbohydrate complexes from straw are usually
extracted with different processes which depend on the lignin
structure and the type of straw. The most current methods for LCC
separation are the hot water and alkaline extractions of ball-milled
straws [17]. The association of phenolic components with carbohydrates presents the greatest barrier to their utilisation. There are
still discussions on whether this barrier is mainly caused by the
lignin (polyphenols), oligomeric phenols or monophenols which
depend on the species, plant part and plant maturity [55].

5. Lignin as adhesives
A primary objective of lignin utilisation research has been to
use industrial lignin as binders; most of the research for this
specic property has focused on the incorporation of lignin with
phenolic wood adhesives for panel products. A detailed report of
the development of lignin-based wood adhesives has been produced by the previous researchers [56]. The thermosetting adhesives used in MDF and PB are mainly urea formaldehyde (UF) but
also melaminureaformaldehyde (MUF) and phenolformaldehyde (PF). The adhesive components are derived from petroleum,
which is increasingly more expensive. A further problem is that
the cured adhesives within the panel products limit the reuse
options of the discarded boards and production residuals [57].
Both economical and health benets could thus be obtained from
binder-less (synthetic resin-free) production processes. One of the
most traditional applications of using lignin as binder is breboard
production, where breboard thermo mechanical pulping (TMP)
converts wood chips to largely lignin-covered bres by shearing
wood bres along the lignin-rich middle lamellae [5861]. The
production of binder-less breboard using enzymatic treated rape

96

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

Fig. 3. Structure model of wheat straw lignin [43].

as probably toxic to humans. The use of formaldehyde-based adhesives also contaminates environment and causes health damages to
workers and consumers [63]. The development of non-toxic adhesives from renewable biomass has signicant interests [6466]. Lignin
acts together with hemicellulose as a perfect natural adhesive for
straw and any other cellulosic materials. The isolated technical lignin
generally are poor binders for wood composites when compared to
conventional resin systems such as PF resins [67], although industrial
soda bagasse lignin was shown to have much higher reactivity toward
formaldehyde, which was considered to be due to a lower degree of
condensation and more phenyl-propanoid units that may be from
p-coumarate [68].
Fig. 4. Lignin phenolic carbohydrate complexes in wheat straw [43].

straw bre has also been reported [62]. In order to re-activate


inactive lignin that is present at the bre surface as well as in the
form of loose linkages between lignin and the other elements,
lignololytic enzymes, such as laccase and peroxidase are to be
added. It is believed that this leads to the formation of functional
groups that are more reactive and an additional bonding in the
material during mat pressing [58]. Comparable mechanical properties to synthetic resin bound breboards can be achieved but
with the long incubation time of the bres, high costs for the
necessary enzymes and high values of thickness-swelling in 24 h
water soaking.
5.1. Optimisation of lignin
As the industry moves towards more sustainable engineering
products, there will be a greater need for sustainable adhesives to
convert these renewable materials into serviceable products. Adhesives currently used for wood-panel products contain formaldehyde
resin which was classied by the World Health Organization (WHO)

5.1.1. Phenolation
A great deal of research work has been carried out to modify
lignin materials. Phenolation is carried out to improve the reactivity position of rice hull acid-insoluble lignin after hydrolysis,
thus making it easy for lignin molecules to incorporate into the
resin through the polymerisation [69]. Ysbrandy et al. modied
auto hydrolysis bagasse lignin through phenolation and used the
phenolated bagasse lignin to make lignin-based PF resin [68].
Physical properties of lignin-based PF resin impregnated paper
laminates were evaluated and the effect of substitution levels of
different phenolic components with lignin was studied. The results
indicated that the lignin-based PF resin with 33% phenolated bagasse
lignin produced laminates with better physical properties [68].
Further phenolated soda bagasse lignin was used to prepare ligninbased PF resin [70,71]. Liu et al. [71] specically phenolated wheat
straw soda lignin and used it to substitute up to 70% of the phenol in
formulated PF resin that was found to have comparable performance
to the original PF resin. The phenolation of rice hull acid-insoluble
lignin was carried out [72] in combination with three step polymerisation of the resin to improve the reactivity of rice hull acid-insoluble

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

lignin, it was found that the good properties of the product, such as
morphology, temperature stability, yields, water and ultraviolet
resistance are due to the increase of the covalent interaction between
the resin and lignin.
5.1.2. Ultraltration
The second approach used for lignin optimisation is ultraltration. Ultraltration is used when lignin of a specic molecular
weight is required for a special product; in a kraft mill, CO2
precipitation and H2SO4 precipitation are the most appropriate
methods [73]. Technical lignins puried by ultraltration process
were found to be promising materials for making lignin-based PF
adhesives and Forss and Fuhrman [73] claimed that adhesives
made with phenolic resin and high molecular weight kraft lignin
fractions illustrate improved adhesive strength. However the
adhesive made from ultraltration-fractionated ammonium-based
sulte spent liquors (SSL) was found to produce high quality
wafer-board with low molecular weight ( o5000) lignosulfonates
while poor mechanical properties wafer-boards with high molecular weight (45000) lignosulfonates [74]. The removal of high
molecular weight lignin by ultraltration was also found to be
feasible and economically attractive in simulation studies [75]. The
two main advantages of using ultraltration for lignin recovery are
that no adjustment of the pH or temperature is needed and the
concentration of the liquor to be treated is not crucial [76].
5.1.3. Biological pre-treatments
Biological pre-treatment is milder in its operational conditions
than traditional physical or chemical pre-treatments. Biological pretreatment has been studied for the production of wood based
composites. The treatment can be divided into two main approaches:
the enzymatic activation of lignin and laccase-assisted adhesion. The
enzymatic activation of lignin for lignocellulosic products such as
MDF and particleboard can improve the self-bonding properties of
the biomass by oxidation of their surface lignin before being
fabricated into boards [77]. The bonding mechanism of breboards
made from laccase treated bres appears to be connected to phenoxy
radicals located on bre surfaces which may cross-link when the
bres are pressed into boards [58,59,7880]. However, signicant
contributions to the adhesive effect by condensation of hemicellulose
degradation products, hydrogen bonding and molecular entanglement cannot be ruled out [79,80]. In the laboratory-scale experimented by Felby et al. [58] and Kharazipour et al. [81,82], bres were
either incubated with laccase [58,79,81] or peroxidase [82] at low
consistency in a water medium at a suitable pH or laccase solution
was sprayed onto the wood bres [81]. For the manufacture of
binder-less wood boards, the laccase-assisted adhesion could have a
great potential provided that the application technology is compatible with existing manufacturing facilities and processes and that
the dimensional stability of the boards can be improved [83].
5.1.4. Reactivity improvement
The fourth approach of optimising the strength properties of
lignin-based adhesives is to increase the reactivity of technical
lignins toward formaldehyde with demethylation [84], hydroxyalkylation [85] and phenolation [86]. Olivares et al. [87] found
that different fractions of softwood lignin separated by ultraltration after methylation and demethylation reacted differently
toward formaldehyde, which then consequently led to different
mechanical and water absorption properties. Yang and Liu [88]
reported that hydroxylmethylation produces resins with enhanced
properties in comparison to original lignin. The increase in lignin
reactivity by reaction with K2Cr2O7, and subsequent formation of
catechol groups through demethylation has been reported previously
[89]. Dichromate in the presence of acetic acid was used [90],

97

decreasing the concentration of methoxyl groups from 11.9 to 2.8%,


and increasing the phenolic hydroxyl groups from 2.4 to 11.2%.
Alternative chemical modication of lignin is methylolation which
is a simple way for increasing lignin reactivity by introducing
methylol groups by reacting lignin with formaldehyde has been
reported [91]. In this study, lignin in alcoholic solution was treated
with formaldehyde (lignin: formaldehyde1: 1) for 12 h at 40 1C and
at pH8, with constant agitation. These conditions were reported as
most favourable by [90] where the kinetics and mechanism of the
reaction of lignin with formaldehyde were studied.
The reactivity of technical lignin toward phenol-oxidising
enzymes is likely to depend on their molecular weight distribution, water-solubility, content of phenolic hydroxyl groups and
number of available phenoxy radical cross-linking sites [92].
None of the chemical modication methods can be used commercially for the production of lignin based products due to the
environmental concerns; however a commercial chemical method
of lignin degradation based on alkaline reactions of lignin is used in
the alkaline pulping process of woods, NaOH serves as the hydrolytic
reagent for the degradation of the lignosulfonates [17].

5.2. Lignin as phenolformaldehyde resins


As already emphasised, the phenolic nature of lignin makes it
prone to replace phenol with lignin derivatives in PF resins to
formulate wood composite adhesives suitable for plywood, particleboard and other similar kind of composites, replacing current
synthetic PF resins which are based on petrochemical, nonrenewable materials from fossil fuel. Lignosulfonates are the most
used technical lignins for making lignin-based adhesives [6],
produced as by-products from sodium, calcium and ammoniumbased sulphite spent liquors (SSL). These lignin derivatives in SSL
can be used directly for formulating wood adhesives [56]. Broad
research has been carried out with a number of different lignin
types as substitutes for phenol in phenolformaldehyde resins.
Feldman [93] has comprehensively reviewed the work in this eld.
The properties of wood adhesive products produced with ligninbased phenolformaldehyde resins have been found to be similar
with those of commercial resins up to 35 wt% partial replacement
with lignin [94]. A variety of different lignins, including organosolv
lignin, soda lignin and lignosulfonates, have been used in phenol
formaldehyde resin preparation [95], also black liquor has been
applied directly [96].
Phenolformaldehyde adhesives have also been prepared from
lignosulfonate resulting from grasses such as bagasse and wheat
straw [71,97]. Liu et al. [98] work indicated that the best adhesive
properties on incorporation into phenolformaldehyde resins of
wheat straw soda-lignin are for the lower molecular weight
fractions, and Peng and Riedl [99] proved in their work that the
reactivity of lignosulfonate with formaldehyde increases when
wheat starch is added as ller, when starch derived from wheat
was used, it produced the lowest level of condensation resulting in
the highest reactivity of the lignin-starch combinations to formaldehyde. Khan et al. [100] assessed the possibility of preparing
wood adhesives from bagasse lignin. They have optimised the
parameters for the preparation of lignin phenolformaldehyde
(LPF), such as lignin concentration, catalyst concentration, reaction
time and temperature, and formaldehyde to phenol molar ratio.
The results indicated that up to 50% of phenol can be substituted
by bagasse lignin to achieve comparable bonding strength when
compared to a commercial phenolformaldehyde wood adhesive.
The optimised condition for preparation of bagasse lignin-based
resin was as follows: lignin to phenol ratio of 50%, formaldehyde to
phenol molar ratio of 2, 10% catalyst concentration of phenol,
reaction time of 4 h and the reaction temperature of 80 1C.

98

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

The use of acetosolv lignins in phenolformaldehyde resin and


the direct replacement of organosolv lignin for phenol in phenol
formaldehyde resins have both been investigated. Plywood made
with the former showed better knife-test shear strength compared
to those obtained with a commercial phenolformaldehyde resin
[101]. Organosolv lignin could replace 2030% of the phenol in PF
resins for particleboards production without adversely effect of
bonding properties [102,103].
Zhang et al. [104] used ethanol residue (ER), which is the byproduct of lignocellulosic ethanol production with rich activated
lignin, to partially replace phenol in the range of 1070% to prepare
ligninphenolformaldehyde (LPF) adhesive. The results revealed
that ER, with rich hydroxyl group and less methoxyl group as well
as lower molecular weight, was suitable for the synthesis of LPF.
In another study by Ibrahim et al. [105] kraft lignin from oil
palm empty fruit bunch showed great potential as a partial
substitute for phenol in PF resin production. El Mansouri et al.
[106], by Mannich reactivity and UV-spectroscopy, found that kraft
and soda-anthraquinone lignin are more reactive towards modication hence good raw material for LPF. Jin et al. [107] synthesised PF resin modied with enzymatic hydrolysis lignin (EHL) by a
one-step process. EHL extracted from residues of cornstalk was
partially substituted by the phenol component in PF resin. Plywood using this resin was produced and the bonding strength
almost met the Chinese National Standard (GB/T 14732-2006).
In summary, there are different methods used to derive lignin
for formation of phenolformaldehyde resins which are as follow:
(1) the lignin reacts with phenol and the ligninphenol complex is
then reacted with formaldehyde, (2) the lignin which reacts with
phenol and formaldehyde, and the pre-polymer is then reacted
with phenol, (3) phenol reacts with formaldehyde and the mixture
is then reacted with lignin, and (4) the lignin reacts with
formaldehyde and the hydroxymethylated lignin is then reacted
with phenol.
The method of extraction and the source of lignin have a strong
inuence on the properties of the phenolformaldehyde resin [108].
The variations by different researchers are consistent throughout
different studies. One of the drawbacks of lignin-based resins
compared to PF resins is that they usually tend to have weaker
adhesion properties and tend to have a high degree of variability in
adhesion performance [109]. Previous studies have shown that the
presence of the plasticisers or contaminants (e.g. very low molecular weight lignin monomeric units) is partially responsible for the
low bond strength [110]. The cost of the technical lignin adhesive
can be signicant; there is also the dilemma that the water-soluble
technical lignin reduces the dimensional stability of the composites,
while water-insoluble lignin may not be reactive enough to provide
sufcient cross-linking.
The lack of optimising lignin for its adhesive properties is very
signicant. The creation of the same environment in the cellulose
materials, in order for lignin to function as it does, with hemicellulose, inside the cellulose materials is what could lead to the
development of bio-based adhesives. In general it has been impractical to use LPF resin as adhesive, due to several reasons, one of
which is related to the molecular size of the adhesive, the lignin
molecules have generally been too large to penetrate the surface of
the wood to get good adhesion also size of lignin molecules prevent
its condensation with phenol to any meaningful extent, hence for
these reasons, when large amounts of lignin are employed in such
resins, adhesive properties have been found to be deteriorate.
There are always reports based on simple substitution of phenol
in PF resin. However this is an old technology and unfortunately not
a successful one in industrial scales. In general the researchers on this
topic do not address the problems of industrialising the PF lignin
adhesives. New and alternative technologies need to be developed
before lignin can be used commercially as a wood adhesive.

5.3. Lignin as epoxy resin


Lignin has also been used in epoxy resins and many different
formulation approaches have been explored [111113]. The synthesis of lignin-based epoxy resins can be divided into three categories as follows: (1) Blending the epoxy resins with technical lignin
obtained directly from the pulp and paper process [114116],
(2) Direct modication of technical lignin by epoxides
[112,117,118], and (3) Modication of lignin by several chemical
reactions in order to improve their reactivity prior to the epoxidation reaction [117,118]. Previous studies have shown that the epoxy
modied lignin-based resins can be prepared in signicant quantities ( 0.5 ML/annum) with relative simple purication schemes
involving the reduction of lignin with epichlorohydrin, subsequent
reclamation of the unused epichlorohydrin and ltration to purify
the ultimate resin [119,120].
A very wide variety of comonomers and curing reagents have
been applied to prepare lignin-derived epoxy resins. The effect of
lignin blending with epoxy resins are strongly affected by the type
of lignin used [121]. Ismail et al. [122] modied sodium lignosulfonate by reaction with anhydrides to form ester-carboxylic acid
derivatives which are used to crosslink glycerol diglycidyl ether
and ethylene glycol diglycidyl ether to obtain the bio-base epoxy
resins. Different methods create different products having diverse
physicochemical properties; hence they may be used in different
applications. Marginally soluble lignin has been formulated with
epoxy resin and cross-linked by heat, the epoxy resins that
contained approximately 50% lignin were successfully prepared
and applied to printed circuit board (PCB) production [123], which
is the most novel development and this resin had physical and
electrical properties similar to those of common laminate resins.
Epoxy resins are known to be one of the most important
thermoset polymers, the market of which is large. With specic
reference to phenolepoxy resins, lignin could ourish as a crosslinking agent [16].

6. Lignin for other applications


The history of lignin utilisation traces back to the late 19th
century when lingosulfonates produced from the bisulte pulping
process were claimed to be effective leather-tanning agents and dyebath additives [124]. Most lignin waste is burned to generate energy
for the pulp mills. Nevertheless, because of its unique properties,
lignin offers perspective for higher added value applications in
renewable products. Technical lignins are main by-product produced
in pulping in the papermaking industry. The water solubility and
reactivity of lignin is increased due to the introduction of hydrophilic
groups and cleavage of linkages between structural units of lignin in
the pulping process. This makes it further chemical modication
possible and many kinds of useful chemicals can be manufactured
from the technical lignins [125]. However, it has always been and still
is very challenging task to use technical lignins efciently and cost
effectively. The major difculties for a wide utilisation of technical
lignin (i.e. kraft lignin, lignosulfonate, soda lignin, organosolv and
ethanol process lignins) relate to the inherent complex nature and
the lack of good understanding or information about their chemical
or physical properties [3]. In addition to the adhesives aforementioned, the focus of utilisation of lignin is on their function as
dispersants, binders and emulsiers for other engineering applications [6]. These lignin types not only lack sulfur groups, but possess
interesting properties for different applications. Different technical
purposes for lignin have been investigated based on its properties
and the modied lignin has been used in cross-linked polyurethanes
or phenolic resin [126]. Conventional kraft and sulphite lignin are
products that have been used as dispersants and binders. However,

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

the most interesting new applications are related to sulfur-free lignin


[127], the reason being for their versatility and processed without
odour release which is commonly observed with commercial kraft
lignin. Among lignin chemical products, lignosulfonate obtained from
spent sulte pulping liquors are the most available. Lignosulfonate
has excellent dispersing properties and has been used as a superplasticiser in concrete or gypsum to improve their uidity.
6.1. Emerging utilisation of straw biomass lignin
With the increased awareness of and environmental pressures to
reduce greenhouse gas emissions, plus other economic motivations,
the power industry has become more interested in opportunities to
cost effectively utilise biomass materials. The greenhouse effect and
the scarcity of petroleum are driving the inventive utilisation of
renewable resources (biomass) on earth. Ensuring the establishment
of adequate, affordable, efcient and reliable high-quality energy
services with minimum or no adverse effect on the environment for
a sustained period is the key for development. Bioethanol production
from lignocellulosic biomass through the hydrolysis of polysaccharides and the subsequent fermentation of monomer sugars is another
important factor for further research on lignin and one of the ways to
improve the energy dependence on fossil oil [128]. Lignin itself can
also be converted into a transportation fuel by dehydroxygenation or
zeolite upgrading [129]. One-step process to convert lignin into nonviscous organic liquid and oil has been developed in 2008 [130].
Shabtai et al. [131] described a two-stage process in their work for
conversion of lignin into high quality reformulated gasoline compositions. Biomass-derived fuels share many of the same characteristics
as their fossil fuel counterparts [132134]. Once formed, they can be
substituted in whole or in part for petroleum-derived products.
As a raw material, lignin can also be used for various value added
products such as vanillin, ferulic acid, coumaric acid, guaiacol,
catechol, HPHTM (hydroxyphenyl-(hydroxytolyl)-methane) and
others [31,135138].
Diversication of energy sources, agricultural activities and a
higher percentage of locally produced energy are goals that can be
satised by biofuels. Biofuels such as bioethanol may be easier to
commercialise than other fuels due to their acceptable performance, low infrastructure investment and other factors. However,
there are some challenges. In addition to polymeric carbohydrates,
plant matter contains varying amounts of polyphenolic lignin and
other extractives. The pre-treatment process which aims to separate the carbohydrates from the lignin matrix while minimising
chemical destruction of fermentation sugars required for ethanol
production is difcult, given that biomass includes sources like
hardwood and softwood trees, agricultural residues like corn
stover and non-recyclable paper waste. Thus different feedstock
have caused researchers to test numerous pre-treatment processes
ranging from hot water and steam explosion treatments, to alkaline and solvent pre-treatments. The comprehensive and successful utilisation of straw biomass remains a difcult task, whatever
the intended application. Overcoming the barriers, which prevent
commercial exploitation of lignocellulose, will be the key to its
successful application in biotechnological accomplishments.
Success with innovative production and technology is critical to
expanding nancial and governmental support for this concept.
The intensity of research and the magnitude of capital investment
on straw biomass commercialisation shall increase immensely
once commercial feasibility seems probable.
7. Conclusions and outlook
It can be gathered from this review that lignin in herbaceous
crops, straw, is a potential source of valuable bio-products.

99

Unfortunately lignin remains largely unutilised due to the lack of


commercially viable production technologies and variation and
complexity of straw lignin. Lignin is a very abundant naturally
occurring polymer with an aromatic and highly cross-linked
structure, similar to the network of phenolformaldehyde resins.
Lignin can be directly used as a macromolecular material without
pre-treatment or modied for specic properties for many materials applications, which can play an important role in fully or
partly replacing petroleum-based components in a broad range of
engineering products. Lignin of superior quality can afford a
signicant opportunity to apply it to a much greater extent in
controlled-release formulations and as a feedstock for fuels and
commodity chemicals. There are technical and economic issues
that need to be addressed before lignin-based engineering products can be produced industrially. However, further studying the
lignin within the straw and other herbaceous crops, especially the
creation of the functionalities of lignin as it does with cellulose
and hemicellulose, could lead to radical development of lignin as
green engineering and biomass as biofuel or other high value
added products for many industrial sectors.
Lignin-based adhesives have potentials for engineering applications due to their environmental suitability and economic and
technical feasibility. Unfortunately researchers have in general not
addressed the nee of their industrialisation. New and advanced
technologies are being developed and new materials for ligninbased adhesives will be created in the future.
References
[1] Morrison I. The degradation and utilization of straw in the rumen. In:
Grossbard E, editor. Straw decay and its effects on disposal and utilization.
Chichester: Wiley; 1979. p. 23745.
[2] Sun R, Tomkinson T. Essential guides for isolation/purication of polysaccharides. In: Woilson I, Adlard T, Poole C, Cook M, editors. Encyclopedia of
separation science. London: Academic Press; 2000. p. 456874.
[3] McCarthy JL, Islam A. Lignin chemistry, technology, and utilization: a brief
history. American Chemical Society; 1999; 299.
[4] Sarkanen K, Ludwig C. Lignins: occurrence, formation, structure and reactions. New York: Wiley; 1971.
[5] Sjstrm E. Wood chemistry: fundamental and applications. 2nd ed. New York:
Academic Press, Inc.; 1993.
[6] Sun R. Lignin. In: Lu F, Ralph J, editors. Cereal straw as a resource for
sustainable biomaterials and biofuels. Amsterdam, the Netherlands: Elsevier;
2010. p. 169207.
[7] Ralph J, Lundquist K, Brunow G, Lu F, Kim H, Schatz P, et al. Lignins: natural
polymers from oxidative coupling of 4-hydroxyphenyl propanoids. Phytochemistry Reviews 2004;3:2960.
[8] Higuchi T. Lignin biochemistry: biosynthesis and biodegradation. Wood
Science and Technology 1990;24:2363.
[9] Lin S, Dence C. Methods in lignin chemistry. Heidelberg: Springer; 1992.
[10] Eriksson Karl-Erik L. Biotechnology in the pulp and paper industry: an
overview. American Chemical Society; 1989; 214.
[11] Kim S, Dale BE. Global potential bioethanol production from wasted crops
and crop residues. Biomass and Bioenergy 2004;26:36175.
[12] Adler E. Lignin chemistry-past, present and future. Wood Science and
Technology 1977;11:169218.
[13] Liu C, Wyman CE. The effect of ow rate of compressed hot water on xylan,
lignin, and total mass removal from corn stover. Industrial & Engineering
Chemistry Research 2003;42:540916.
[14] Wool R, Sun X. Polymers and composites. In: Wool R, editor. Bio-based
polymers and composites. California: Elsevier; 2005. p. 55198.
[15] Thielemans W, Can E, Morye SS, Wool RP. Novel applications of lignin in
composite materials. Journal of Applied Polymer Science 2002;83:32331.
[16] Stewart D. Lignin as a base material for materials applications: chemistry,
application and economics. Industrial Crops and Products 2008;27:2027.
[17] Buranov AU, Mazza G. Lignin in straw of herbaceous crops. Industrial Crops
and Products 2008;28:23759.
[18] Sahoo S, Seydibeyolu M, Mohanty AK, Misra M. Characterization of
industrial lignins for their utilization in future value added applications.
Biomass and Bioenergy 2011;35:42307.
[19] Beckman E, Liesche O. Qualitative und quantitative unterschiede der lignine
einiger holz-undstroharten. Biochemische Zeitschrift 1923;139:491508.
[20] Jackson MG. Review article: the alkali treatment of straws. Animal Feed
Science and Technology 1977;2:10530.
[21] Ryan K. Comparison between utilization of cellulose for paper from wood
and straw. In: Hill R, Munck L, editors. New approaches to research on cereal
carbyhydrates. Amsterdam: Elsevier; 1985. p. 3237.

100

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

[22] Scalbert A, Monties B, Rolando C, Sierra-Escudero A. Formation of ether


linkage between phenolic acids and graminear lignin: a possible mechanism
involving unique methods. Holzforschung 1986;40:1915.
[23] Scalbert A, Monties B, Lallemand J, Guittet E, Rolando C. Ether linkage
between phenolic acids and lignin fractions from wheat straw. Phytochemistry 1985;24:135962.
[24] Eraso F, Hartley RD. Monomeric and dimeric phenolic constituents of plant
cell walls? Possible factors inuencing wall biodegradability Journal of the
Science of Food and Agriculture 1990;51:16370.
[25] Bjrkman A. Lignin and lignincarbohydrate complexes. Industrial & Engineering Chemistry 1957;49:13958.
[26] Ralph J, Helm RF, Quideau S, Hateld RD. Lignin-feruloyl ester cross-links in
grasses. Part 1. Incorporation of feruloyl esters into coniferyl alcohol
dehydrogenation polymers. Journal of the Chemical Society, Perkin Transactions 1 1992;0:29619.
[27] Merewether J. A lignincarbohydrate complex in wood. Holzforschung
1957;11:6580.
[28] Lapierre C, Pollet B, Rolando C. New insights into the molecular architecture
of hardwood lignins by chemical degradative methods. Research on Chemical Intermediates 1995;21:397412.
[29] Lapierre C. Application of new methods for the investigation of lignin
structure. In: Jung H, Buxton D, Hateld R, Ralph J, editors. Forage cell wall
structure and digestibility. Madison: American Society of Agronomy; 1993.
p. 13366.
[30] Zilliox C, Debeire P. Hydrolysis of wheat straw by a thermostable endoxylanase: adsorption and kinetic studies. Enzyme and Microbial Technology
1998;22:5863.
[31] Tapin S, Sigoillot J, Asther M, Petit-Conil M. Feruloyl esterase utilization for
simultaneous processing of nonwood plants into phenolic compounds and
pulp bers. Journal of Agricultural and Food Chemistry 2006;54:3697703.
[32] Azuma J, Tetsuo K. Lignincarbohydrate complexes from various sources.
Methods in enzymology. Academic Press; 1988; 128.
[33] Billa E, Koukios EG, Monties B. Investigation of lignins structure in cereal
crops by chemical degradation methods. Polymer Degradation and Stability
1998;59:715.
[34] Schultz TP, Fisher TH, Dershem SM. Role of the p-hydroxyl group in the
nitrobenzene oxidation of hydroxybenzyl alcohols. Journal of Organic
Chemistry 1987;52:27981.
[35] Galletti GC, Piccaglia R, Chiavari G, Concialini V. HPLC/electrochemical
detection of lignin phenolics from wheat straw by direct injection of
nitrobenzene hydrolysates. Journal of Agricultural and Food Chemistry
1989;37:9857.
[36] Sun RC, Fang JM, Tomkinson J. Delignication of rye straw using hydrogen
peroxide. Industrial Crops and Products 2000;12:7183.
[37] Del Ro J, Gutierrez A, Rodriguez I, Ibarra D, Martinez A. Composition of nonwoody plant lignins and cinnamic acids by Py-GC/MS, Py/TMAH and FT-IR.
Industrial Crops and Products 2007;79:3946.
[38] Morrison I. A semi-micro method for the determination of lignin and its use
inpredicting the digestibility of forages. Journal of the Science of Food and
Agriculture 1972;23:45563.
[39] Van Soest P, Wine R. Determination of lignin and cellulose in acid-detergent
bre with permanganate. Journal of the Association of Ofcial Analytical
Chemists 1968;51:7805.
[40] Fidalgo M, Terron M, Martinez A, Gonzalez A, Gonzalez-vila F, Galletti G.
Comparative study of fractions from alkaline extraction of wheat straw
through chemical degradation, analytical pyrolysis, and spectroscopic techniques. Journal of Agricultural and Food Chemistry 1993;41:16216.
[41] Jung HJG, Himmelsbach DS. Isolation and characterization of wheat straw
lignin. Journal of Agricultural and Food Chemistry 1989;37:817.
[42] Scalbert A, Monties B, Guittet E, Lallemand J. Comparison of wheat straw
lignin preparations I. Chemical and spectroscopic characterizations. Holzforschung 1986;40:11929.
[43] Sun R, Lawther JM, Banks WB. A tentative chemical structure of wheat straw
lignin. Industrial Crops and Products 1997;6:18.
[44] Scalbert A, Monties B, Lallemand J, Guittet E, Rolando C. Ether linkage
between phenolic acids and lignin fractions from wheat straw. Phytochemistry 1985;24:135962.
[45] Sun R. Structure, ultrastructure, and chemical composition. In: Xu F, editor.
Cereal straw as a resource for sustainable biomaterials and biofuels.
Amsterdam, the Netherlands: Elsevier; 2010. p. 947.
[46] Baucher M, Monties B, Montagu MV, Boerjan W. Biosynthesis and genetic
engineering of lignin. Critical Reviews in Plant Sciences 1998;17:12597.
[47] Donaldson LA. Lignication and lignin topochemistryan ultrastructural
view. Phytochemistry 2001;57:85973.
[48] Saka S, Goring D. Localization of lignins in wood cell walls. In: Higuchi T,
editor. Biosynthesis and biodegradation of wood components. Orlando:
Academic Press; 1985. p. 5162.
[49] Terashima N, Fukuzuka T, He L, Takabe K. Comprehensive model of the
lignied plant cell wall. In: Jung H, Buxton D, Hateld R, Ralph J, editors.
Forage cell wall structure and digestibility. Madison: American Society of
Agronomy; 1993. p. 24770.
[50] Chang H, Cowling E, Brown W, Adler E, Miksche G. Comparative studies on
cellulolytic enzyme lignin and milled wood lignin of sweetgum and spruce.
Holzforschung 1975;29:1539.
[51] Bjrkman A. Studies on nely divided wood. Part I. Extraction of lignin with
neutral solvents. Sven Papperstidn 1956;59:47785.

[52] Sun R, Lawther JM, Banks WB. The effect of alkaline nitrobenzene oxidation
conditions on the yield and components of phenolic monomers in wheat
straw lignin and compared to cupric(II) oxidation. Industrial Crops and
Products 1995;4:24154.
[53] Lawther JM, Sun R. The fractional characterisation of polysaccharides and
lignin components in alkaline treated and atmospheric rened wheat straw.
Industrial Crops and Products 1996;5:8795.
[54] Sun R, Tomkinson J. Comparative study of lignins isolated by alkali and
ultrasound-assisted alkali extractions from wheat straw. Ultrasonics Sonochemistry 2002;9:8593.
[55] Himmelsbach DS. Structure of forage cell walls. In: Jung H, Buxton D,
Harteld R, Ralph J, editors. Forage cell wall structure and digestibility.
Madison: American Society of Agronomy; 1993. p. 27183.
[56] Nimz H. Lignin-based wood adhesives. In: Pizzi A, editor. Wood adhesive:
chemistry and technology. New York: Marcel Dekker; 1983. p. 24788.
[57] Smith D. The generation and utilization of residuals from composite panel
products. Forest Products Journal 2004;25:817.
[58] Felby C, Pedersen L, Nielsen B. Enhanced auto-adhesion of wood bres using
phenol oxidases. Holzforschung 1997;51:2816.
[59] Kharazipour A, Huettermann A, Luedemann HD. Enzymatic activation of
wood bres as a means for the production of wood composites. Journal of
Adhesion Science and Technology 1997;11:41927.
[60] Widsten P, Laine J, Qvintus-Leino P, Tuominen S. Effect of high temperature
brization on the chemical structure of hardwood. Holzforschung 2002;56:
519.
[61] Widsten P, Laine J, Qvintus-Leino P, Tuominen S. Effect of high temperature
brization on the chemical structure of softwood. Journal of Wood Chemistry and Technology 2001;21:22745.
[62] Unbehaun H, Dittler B, Khne G, Wagenfhr A. Investigation into the
biotechnological modication of wood and its application in the woodbased material industry. Acta Biotechnologica 2000;20:30512.
[63] Lin Q, Chen N, Bian L, Fan M. Development and mechanism characterization
of high performance soy-based bio-adhesives. International Journal of
Adhesion and Adhesives 2012;34:116.
[64] Pizzi A. Recent developments in eco-efcient bio-based adhesives for wood
bonding: opportunities and issues. Journal of Adhesion Science and Technology 2006;20:82946.
[65] Liu Y, Li K. Development and characterization of adhesives from soy protein
for bonding wood. International Journal of Adhesion and Adhesives
2007;27:5967.
[66] Despres A, Pizzi A, Vu C, Pasch H. Formaldehyde-free amino resin wood
adhesive based on dimethoxyethanal. Journal of Applied Polymer Science
2008;110:390816.
[67] Lewis N, Lantzy T, Branhm S. Lignin in adhesives: introduction and historical
perspective. In: Hemingway R, Conner A, editors. Adhesives from renewable
resources. Oxford: Oxford University Press; 1989. p. 1326.
[68] Ysbrandy R, Sanderson R, Gerischer G. Adhesives from autohydrolysis
bagasse lignin, a renewable resourcePart II. DSC thermal analysis of
novolac resins. Holzforschung 1992;46:2536.
[69] Gosselink RJA, Abcherli A, Semke H, Malherbe R, Kuper P, Nadif A, et al.
Analytical protocols for characterisation of sulphur-free lignin. Industrial
Crops and Products 2004;19:27181.
[70] Vzquez G, Gonzlez J, Freire S, Antorrena G. Effect of chemical modication
of lignin on the gluebond performance of ligninphenolic resins. Bioresource
Technology 1997;60:1918.
[71] Liu G, Qiu X, Xing D, Yang D. Phenolation modication of wheat straw soda
lignin and its utilization in lignin-based formaldehyde resins. In: He B, Fu S,
Chen F, editors. In: Third international symposium on emerging technologies
of pulping and papermaking, Guangzhou: South China University Technology Press; 2006. p. 933938.
[72] Ma Y, Zhao X, Chen X, Wang Z. An approach to improve the application of
acid-insoluble lignin from rice hull in phenolformaldehyde resin.
Colloids and Surfaces A: Physicochemical and Engineering Aspects 2011;377:
2849.
[73] Forss K, Fuhrmann A. Finnish plywood, particleboard, and breboard made
with a lignin-base adhesive. Forest Products Journal 1979;29:259.
[74] Shen K, Calve L. Improving by fractionation-ammonium-based spent sulphite
liquor for waferboard binder. Adhesives Age 1980;23:259.
[75] Kirkman A, Gratzl J, Edwards L. Kraft lignin recovery by ultraltration:
economic feasibility and impact on the kraft recovery system. Tappi Journal
1986;69:1104.
[76] Jnsson A, Wallberg O. Cost estimates of kraft lignin recovery by ultraltration. Desalination 2009;237:25467.
[77] Widsten P, Kandelbauer A. Adhesion improvement of lignocellulosic products
by enzymatic pre-treatment. Biotechnology Advances 2008;26:37986.
[78] Felby C, Hassingboe J, Lund M. Pilot-scale production of berboards made by
laccase oxidized wood bers: board properties and evidence for crosslinking of lignin. Enzyme and Microbial Technology 2002;31:73641.
[79] Felby C, Thygesen LG, Sanadi A, Barsberg S. Native lignin for bonding of ber
boardsevaluation of bonding mechanisms in boards made from laccasetreated bers of beech (Fagus sylvatica). Industrial Crops and Products
2004;20:1819.
[80] Widsten P, Laine JE, Tuominen S, Qvintus-Leino P. Effect of high debration
temperature on the properties of medium-density berboard (MDF) made
from laccase-treated hardwood bers. Journal of Adhesion Science and
Technology 2003;17:6778.

S.H. Ghaffar, M. Fan / International Journal of Adhesion & Adhesives 48 (2014) 92101

[81] Kharazipour A, Huettermann A, Luedemann HD. Enzymatic activation of


wood bres as a means for the production of wood composites. Journal of
Adhesion Science and Technology 1997;11:41927.
[82] Kharazipour A, Bergmann K, Nonninger K, Httermann A. Properties of bre
boards obtained by activation of the middle lamella lignin of wood bres
with peroxidase and H2O2 before conventional pressing. Journal of Adhesion
Science and Technology 1998;12:104553.
[83] Widsten P, Kandelbauer A. Laccase applications in the forest products
industry: a review. Enzyme and Microbial Technology 2008;42:293307.
[84] Gupta R, Singh S, Jolly S. Phenolligninformaldehyde adhesives for plywood. Holzforschung Und Holzverwertung 1978;30:10912.
[85] Peng W, Barry AO, Riedl B. Characterization of methylolated lignin by H-Nmr
and 13C NMR. Journal of Wood Chemistry and Technology 1992;12:299312.

[86] Alonso MV, Oliet M, Rodrguez


F, Garca J, Gilarranz MA, Rodrguez
JJ.
Modication of ammonium lignosulfonate by phenolation for use in phenolic resins. Bioresource Technology 2005;96:10138.
[87] Olivares M, Guzmn JA, Natho A, Saavedra A. Kraft lignin utilization in
adhesives. Wood Science and Technology 1988;22:15765.
[88] Yang S, Liu Q. Effect of lignin modication on synthetic properties. In: Second
international symposium on emerging technologies of pulping & paper
papermaking; 2002. p. 782.
[89] Hayashi A, Namura Y, Urkita T. Demethylation of lignosulphonate during the
gelling reaction with dichromate. Mokuzai Gakkaishi 1967;13:1947.
[90] Jolly S, Singh S, Singh S, Gupta R. Kinetics and mechanisms of lignin
formaldehyde resinication reaction. Cellulose Chemistry and Technology
1982;16:51122.
[91] Marton A. In: Gould Robert F, editor. Lignin structure and reactions, advances
in chemistry series. American Chemical Society; 1966. p. iiv.
[92] Httermann A, Milstein O, Nicklas B, Trojanowski J, Haars A, Kharazipour A.
Enzymatic modication of lignin for technical use. American Chemical
Society; 1989; 36170.
[93] Feldman D. In: Hu T, editor. Lignin and its polyblendsa review. US:
Springer; 2002. p. 8199.
[94] Kulshreshtha AK, Vasile C. Handbook of polymer blends and composites, vol.
14.
[95] Nada A, Abou-Youssef H, El-Gohary S. Phenol formaldehyde resin modication
with lignin. PolymerPlastics Technology and Engineering 2003;42:68999.
[96] Wang Y, Peng W, Chai L, Peng B, Min X, He D. Preparation of adhesive for
bamboo plywood using concentrated papermaking black liquor directly.
Journal of Central South University of Technology 2006;13:537.
[97] Akhtar T, Lutfullah G, Zahoorullah. Lignosulfonate-phenolformaldehyrde
adhesive: a potential binder for wood panel industries. Journal of Chemical
Society of Pakistan 2011;33:5358.
[98] Liu G, Qiu X, Yang D. Properties of wheat straw soda lignin of different
molecular weights and its inuence on properties of LPF adhesive. Huagong
Xuebao (Chinese Edition) 2008;59:15904.
[99] Peng W, Riedl B. The chemorheology of phenolformaldehyde thermoset
resin and mixtures of the resin with lignin llers. Polymer 1994;35:12806.
[100] Khan MA, Ashraf SM, Malhotra VP. Development and characterization of a
wood adhesive using bagasse lignin. International Journal of Adhesion and
Adhesives 2004;24:48593.

[101] Vzquez G, Rodrguez-Bona


C, Freire S, Gonzlez-lvarez J, Antorrena G.
Acetosolv pine lignin as copolymer in resins for manufacture of exterior
grade plywoods. Bioresource Technology 1999;70:20914.
[102] etin NS, zmen N. Use of organosolv lignin in phenolformaldehyde resins
for particleboard production: I. Organosolv lignin modied resins. International Journal of Adhesion and Adhesives 2002;22:47780.
[103] etin NS, zmen N. Use of organosolv lignin in phenolformaldehyde resins
for particleboard production: II. Particleboard production and properties.
International Journal of Adhesion and Adhesives 2002;22:4816.
[104] Zhang W, Ma Y, Xu Y, Wang C, Chu F. Lignocellulosic ethanol residue-based
ligninphenolformaldehyde resin adhesive. International Journal of Adhesion and Adhesives 2013;40:118.
[105] Mohamad Ibrahim MN, Zakaria N, Sipaut CS, Sulaiman O, Hashim R. Chemical
and thermal properties of lignins from oil palm biomass as a substitute for
phenol in a phenol formaldehyde resin production. Carbohydrate Polymers
2011;86:1129.
[106] Mansouri NE, Salvad J. Structural characterization of technical lignins for
the production of adhesives: application to lignosulfonate, kraft, sodaanthraquinone, organosolv and ethanol process lignins. Industrial Crops
and Products 2006;24:816.
[107] Jin Y, Cheng X, Zheng Z. Preparation and characterization of phenol
formaldehyde adhesives modied with enzymatic hydrolysis lignin. Bioresource Technology 2010;101:20468.
[108] Doherty WOS, Mousavioun P, Fellows CM. Value-adding to cellulosic ethanol:
lignin polymers. Industrial Crops and Products 2011;33:25976.

101

[109] Cyr N, Ritchie R. Estimating the adhesive quality of lignins for internal bond
strength. In: Glasser W, Sarkanen S, editors. Lignin: properties and materials.
Washington, DC: ACS; 1989 [Chapter 28].
[110] Hiro-kuni O, Kenichi S. Wood adhesives from phenolysis lignin. In: Glasser W,
Sarkanen S, editors. Lignin: properties and materials. Washington, DC: ACS;
1989 (Chapter 25).
[111] Nieh W, Glasser W. Lignin derivatives with epoxy functionality. In: Glasser
W, Sarkanen S, editors. Lignin: properties and materials. Washington, DC:
ACS; 1989. p. 50614.
[112] Hofmann K, Glasser WG. Engineering plastics from lignin. 22. Cure of lignin
based epoxy resins. The Journal of Adhesion 1993;40:22941.
[113] Hofmann K, Glasser W. Engineering plastics from lignin, 23. Network
formation of lignin-based epoxy resins. Macromolecular Chemistry and
Physics 1994;195:6580.
[114] Feldman D, Banu D, Luchian C, Wang J. Epoxy lignin polyblends: correlation
between polymer interaction and curing temperature. Journal of Applied
Polymer Science 1991;42:130718.
[115] Feldman D, Banu D, Natansohn A, Wang J. Structure properties relations of
thermally cured epoxy lignin polyblends. Journal of Applied Polymer Science
1991;42:153750.
[116] Nonaka Y, Tomita B, Hatano Y. Synthesis of lignin/epoxy resins in aqueous
systems and their properties. Holzforschung 1997;51:1837.
[117] Simionescu C, Rusan V, Turta K, Bobcova S, Macoveanu M, Cazacu G, et al.
Synthesis and characterization of some iron lignosulfonate-based ligninepoxy resins. Cellulose Chemistry and Technology 1993;27:62744.
[118] Malutan T, Nicu R, Popa V. Lignin modication by epoxidation. BioResources
2008;3:13719.
[119] Simionescu CI, Rusan V, Macoveanu MM, Cazacu G, Lipsa R, Vasile C, et al.
Lignin/epoxy composites. Composites Science and Technology 1993;48:31723.
[120] Cazacu G, Popa VI. Lignin-based blends. In: Vasile C, Kulshreshtha A, editors.
Hanbook of polymer blends and composites, vol. 4Bvol. 4B. Shawbury, UK:
Rapra Technology Ltd; 2003. p. 565614.
[121] Muller PC, Glassert WG. Engineering plastics from lignin. VIII. Phenolic resin
prepolymer synthesis and analysis. The Journal of Adhesion 1984;17:15773.
[122] Ismail TNMT Hassan HA, Hirose S, Taguchi Y, Hatakeyama T, Hatakeyama H.
Synthesis and thermal properties of ester-type crosslinked epoxy resins derived
from lignosulfonate and glycerol. Polymer International 2010;59:1816.
[123] Kosbar LL, Gelorme JD, Japp RM, Fotorny WT. Introducing biobased materials
into the electronics industry: developing a lignin-based resin for printed
wiring boards. Journal of Industrial Ecology 2000;4:93106.
[124] LIN SY. Lignin utilization: potential and challenge. Progress in Biomass
Conversion 1983;4:3178.
[125] Chen F, Li J. Aqueous gel permeation chromatographic methods for technical
lignins. Journal of Wood Chemistry and Technology 2000;20:26576.
[126] Kuo M, Hse C, Huang D. Alkali-treated kraft lignin as a component in
akeboard resins. Holzforschung 2009;45:4754.
[127] Lora J, Glasser W. Recent industrial applications of lignin: a sustainable
alternative to nonrenewable materials. Journal of Polymers and the Environment 2002;10:3948.
[128] Woodson M, Jablonowski CJ. An economic assessment of traditional and
cellulosic ethanol technologies. Energy Sources, Part B: Economics, Planning,
and Policy 2008;3:37283.
[129] Piskorz J, Majerski P, Radlein D, Scott DS. Conversion of lignins to hydrocarbon fuels. Energy Fuels 1989;3:7236.
[130] Gellerstedt G, Li J, Eide I, Kleinert M, Barth T. Chemical structures present in
biofuel obtained from lignin. Energy Fuels 2008;22:42404.
[131] Shabtai J, Zmierczak W, Chornet E. Process for conversion of lignin to
reformulated hydrocarbon gasoline. US patent; 1998. 5,959,167.
[132] Klass D. Biomass for renewable energy, fuels and chemicals. New York:
Academic Press; 1988.
[133] Lynd LR. Overview and evaluation of fuel ethanol from cellulosic biomass:
technology, economics, the environment, and policy. Annual Review of
Energy and the Environment 1996;21:40365.
[134] Ma F, Hanna MA. Biodiesel production: a review. Bioresource Technology
1999;70:115.
[135] Pecina R, Burtscher P, Bonn G, Bobleter O. GCMS and HPLC analyses of lignin
degradation products in biomass hydrolyzates. Fresenius' Zeitschrift fr
analytische Chemie 1986;325:4615.
[136] Bobleter O. Hydrothermal degradation of polymers derived from plants.
Progress in Polymer Science 1994;19:797841.
[137] Yoshida T, Matsumura Y. Gasication of cellulose, xylan, and lignin mixtures
in supercritical water. Industrial & Engineering Chemistry Research
2001;40:546974.
[138] Okuda K, Man X, Umetsu M, Takami S, Adschiri T. Efcient conversion of
lignin into single chemical species by solvothermal reaction in water
p-cresol solvent. Journal of Physics: Condensed Matter 2004;16:132530.

You might also like