Modification of Lignin.2002-Leer PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 57

PLEASE SCROLL DOWN FOR ARTICLE

This article was downloaded by:


On: 26 January 2011
Access details: Access Details: Free Access
Publisher Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-
41 Mortimer Street, London W1T 3JH, UK
Polymer Reviews
Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t713597276
MODIFICATION OF LIGNIN
*
John J. Meister
a
a
Forest Products Research Center, Albuquerque, NM, U.S.A.
Online publication date: 24 June 2002
To cite this Article Meister, John J.(2002) 'MODIFICATION OF LIGNIN*', Polymer Reviews, 42: 2, 235 289
To link to this Article: DOI: 10.1081/MC-120004764
URL: http://dx.doi.org/10.1081/MC-120004764
Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf
This article may be used for research, teaching and private study purposes. Any substantial or
systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or
distribution in any form to anyone is expressly forbidden.
The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss,
actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly
or indirectly in connection with or arising out of the use of this material.
MODIFICATION OF LIGNIN*
John J. Meister
Forest Products Research Center, 2008 Hendola Dr., NE,
Albuquerque, New Mexico, 87110-4808
E-mail: polysurlabs@cnsp.com
CONTENTS
I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
A. The Nature of Lignin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
B. Recovery of Lignin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
C. Uses of Extracted Lignin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
II. Modication of Lignin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
A. Decomposition of Lignin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
B. Using Lignin as Extracted from the Plant. . . . . . . . . . . . . . . . . . . 245
C. Adding to the Lignin Biopolymer . . . . . . . . . . . . . . . . . . . . . . . . 251
III. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
I. INTRODUCTION
Lignin [8068-00-6] is a natural product produced by all woody plants. It
is second only to cellulose in mass of the natural polymer formed per
annum.
[1]
Lignin constitutes between 15 and 40 percent of the dry weight
of wood with variation in lignin content being caused by species type,
235
Copyright
#
2002 by Marcel Dekker, Inc. www.dekker.com
*Reprinted from Polymer Modication: Principles, Techniques, and Applications; Meister, J. J.,
Ed.; Marcel Dekker, Inc.: New York; 2000, 67144.
J. MACROMOL. SCI.POLYMER REVIEWS, C42(2), 235289 (2002)
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
growing conditions, the parts of the plant tested, and numerous other fac-
tors.
[2]
The data of Table 1 show the variation of lignin content by species
type. Plants use lignin to (1) add strength and structure to their cellular
composites; (2) control uid ow; (3) protect against attack by micro-
organisms; (4) act as an antioxidant, a UV absorber, and possibly a ame
retardant; and (5) store energy.
[3]
When considering the present and future
use of this biopolymer, it is important to realize that any archeological age,
such as the iron age, starts and also ends before the participants realize it.
We are currently at the end of the age of oil. The slow decline in available oil
reserves during the early 21st century will make lignin a more important
236 MEISTER
Table 1. Lignin Content of U.S. Woods as Determined at U.S. Forest
Products Laboratory from 1927 to 1968
Scientic Name/Common Name Klason Lignin
A. Hardwoods
Acer macrophyllum Pursh/Bigleaf maple 25
Betula alleghaniensis Britton/Yellow birch 21 (2)
a
Carya Cordiformus (Wangenh.)
K. Koch/Bitternut hickory
25
Populus tremoides Michx./Quaking aspen 19 (22)
Quercus falcata Michx./Southern red oak 25
Quercus rubra L./Northern red oak 24
Fagus Grandifolia Ehrh./American beech 22 (2)
Gleditsia tricanthos L./Honey locust 21
Liriodendron tulipifera L./Yellow-poplar 20
Populus deletoides Bartr. Ex Marsh./
Eastern Cottonwood
23 (3)
Salix nigra March./Black willow 21 (2)
B. Softwoods
Abies balsamea (L.) Mill./Balsam r 29 (16)
Larix occidentalis Nutt./Western larch 27 (3)
Picea glauca (Moench) Voss./White spruce 29 (8)
Pinus banksiana Lamb./Jack pine 27 (27)
Pinus elliottii Engelm./Slash pine 27 (15)
Pinus strobus L./Eastern white pine 27 (5)
Sequoia sempervirens (D. Don) Endl./Redwood
Old growth 33
Second growth 33
Tsuga canadensis (L.)
Carr./Eastern hemlock
33 (7)
a
Numbers in parenthesis are number of independent determinations for the
component. In some cases, the trees are from dierent locations. Values are
weight percent contained in moisture-free wood. Data are from Table 3,
p. 76, Chapter 2, Chemical Composition of Wood, By R.C. Pettersen, in
The Chemistry of Solid Wood by R. Rowell, Ed., Advances in Chemistry
Series, Vol. 207, Amer. Chem. Soc., 1984, ISBN 0-8412-0796-8.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
source of chemicals for our future society. When fundamental technology
within a society changes, decades of work preceding the change must have
occurred to develop new technologies to replace those that are obsolete. As
the age of oil changes to the age of biomass, some of the chemical modica-
tions described below will become important industrial processes for pro-
ducing the chemicals and materials that society needs.
A. The Nature of Lignin
Woody plants synthesize lignin from 3-(4-hydroxyphenyl)-2-propenol
(trans-4-coumaryl alcohol 1.1, grasses), 3-(4-hydroxy-3-methoxyphenyl)-
2-propenol (trans-coniferyl alcohol 1.2, pines), and 3-(4-hydroxy-3,5-
dimethoxyphenyl)-2-propenol (trans-sinapyl alcohol 1.3, deciduous) by
free radical crosslinking initiated by enzymatic dehydrogenation.
[4]
Structures of these alcohols and the notation for the carbon atoms of the
C
9
repeat unit of lignin are given in Figure 1. As indicated in the naming of
the alcohols, each class of plants, grasses, softwoods, and hardwoods, pro-
duces a lignin rich in one type of alcohol repeat unit. That lignin is produced
by free radical reaction of the alcohol mixture induced by enzymatic dehy-
drogenation of a C
9
alcohol.
[5]
MODIFICATION OF LIGNIN 237
Figure 1. Structures of the alcohols that form lignin and the notation for the carbon atoms in
the monomers.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
The polymerization of the alkene-substituted phenols can produce a
number of bond structures by delocalization of and reaction at, the free
radical site.
[6]
The lignin produced by a plant is a species and plant-part
specic compound that has dierent composition and structure even within
the same plant. This means that the lignin recovered from a woody plant will
be a mixture of structures and repeat unit compositions that will vary with
the source wood, species, and growing pattern of the wood.
[7]
Lignin then,
identies a class of C
9
-repeat unit, alkylphenol, network polymers formed
with the repeat unit bondings shown in Table 2, where R
1
and R
2
are hydro-
gen or methoxyl groups. The location of the bond between the repeat units is
specied by listing the carbon atom label or heteroatom element symbol for
each atom encountered while moving from one repeat unit to the next.
The most common bond in lignin, the bO4, is a bond starting at the
middle carbon atom (b) of the propyl sidechain on one repeat unit, linking
through the oxygen of the next repeat unit to the number 4, carbon atom of
the aromatic ring of that repeat unit. One structure, g
2
of Table 2, violates the
notation pattern usually used for these bonds but this label for the bi(cyclic
ether) structure is not common. This knowledge of the frequency of these
bonding structures in natural and synthetic lignin is based on assays and
calculations by several authors. The results of these analyses
[811]
are sum-
marized in Table 3. Adler has recalculated the data of Table 3 to express it as
percent of all repeat unit bonds that are of a given type but all of this data is
subject to error introduced by extraction method, processing of the lignin,
and the digestion of the lignin to monomers. Despite the limitations of the
data, it does show clearly that hardwood (Beech) and softwood (Spruce)
lignin dier in bonding structure.
As the lignin monomers react, structures and functional groups not
present in the original alcohols are formed. A tabulation of functional
groups found in milled wood lignin and kraft pine lignin is given in Table 4.
The three-dimensional networks forming lignin are distributed in and
between the plant cells.
The number of repeat units bound together in a lignin agglomeration in
a plant would give a measure of the molecular weight of natural lignin. This
number tells how many aromatic alcohol units are connected to one another,
with the only other bonds outside the network being bonds to pectins, hemi-
cellulose, or cellulose residues. The best value for the lower limit for this
number in softwoods is approximately 60 with a weight average molecular
weight
[13]
for the network fragment of 11,000. This value is probably low
since the milling process used to recover the lignin may have broken bonds in
the network and formed lower molecular weight oligomers of the polyaro-
matic. The actual structure of the three-dimensional network that is the lignin
molecule in the plant has long been assumed
[14]
to be a randomly linked, C
9
lattice as would be produced by a free radical reaction of a resonance-
stabilized, aromatic alcohol.
[15,16]
However, ultraviolet microprobe analysis
238 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
MODIFICATION OF LIGNIN 239
Table 2. Repeat Unit Bondings in Lignin
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
240 MEISTER
Table 4. Functional Groups Found in 100, C
9
Repeat Units of Lignin
Functional Group
Spruce, Milled
Wood Lignin
Pine Lignin,
Kraft Process
Hydroxyl, OH
Total: 120 120
#4-phenolic 30 60
1,2-benzenediol 12
alkyl OH
a
90 48
Carboxylic acid, CO
2
H 5 16
Aldehyde, CO
Total: 20 15
a-propyl 7 5
b-propyl
a
10 10
g-propyl 3
Phenylmethanol and ether
Noncyclic 42 <6
Cyclic
oxacyclopentene
b
11 3
bb, bi(cyclic ether)
c
10 5
Ethene double bond, >CC<
a-phenyl-b, g-ethene 7
d
1,2-diphenylethene 7
Other
d
a
Calculated by dierence.
b
These ring ethers are also called coumaran or benzofuran structures.
[12]
c
See structure g
2
of Table 2.
d
Detected, not quantied.
Table 3. Number of Dierent Bonding Linkages Between 100
Lignin Repeat Units
Type of
Lignin
Loblolly
Pine, MWL
Spruce,
Oxidation
Beech,
Thioacetolysis
Bond
bO4 55 4951 65
aO4 68
b5 16 915 6
b1 9 2 15
55 9 9.5 2.3
4O4 3 3.5 1.5
bb 2 2 5.5
bb* 2
a/gOg 10
ab 11 2.5
b6, 65 2 4.55
a
0
a
1O4, 15 1O4, not seen
a
Combined b6, 65, 1O4, and 15 content.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
of loblolly pine (Pinus taeda L.) thin sections indicates that the aromatic
rings of sizable portions of the lignin are preferentially oriented in the
plane of the cell wall.
[17,18]
These observations would suggest that lignin is
not randomly crosslinked into a naturally occurring, phenolic adhesive (See
Chapter 12) but is built up as an organized, binding structure in the plant.
This view of lignin as a aromatic alcohol truss is supported by the presence of
vesicules
[1921]
in zones of lignication in the cell. A vesicule could allow the
preparation of structured lignin in a specic sequence of enzymatically
mediated steps. Further work is needed to conrm if lignin is an ordered,
network polymer in the plant; determine what fraction of the lignin is ordered
in the loblolly pine; and conrm if this order is common in the rest of the
plant kingdom.
B. Recovery of Lignin
For lignin to be used as the class of chemicals it is, it must be removed
from the plant. Added to the diversity of repeat units and bonding patterns
which characterize natural lignin is the chemical alteration introduced by
each means of removing lignin from wood. Lignin recovery processes
which extract lignin from wood change the chemical and functional group
composition of lignin and make this material extremely heterogeneous.
Methods for recovering lignin are the alkali process, the sulte process,
ball milling, enzymatic release, hydrochloric acid digestion, and organic sol-
vent extraction. Alkali lignins are produced by the kraft
[22]
and soda
[23]
methods for wood pulping. These processes are based on sodium sulfate
plus sodium hydroxide or just sodium hydroxide, respectively. They have
low sulfur content (<1.6 wt.%), contain sulfur contamination present as
thioether linkages, and are water-insoluble, nonionic polymers of 2,000 to
15,000, molecular weight. Over 20 million megagrams of kraft lignin are
produced in the United States each year.
[4]
The sulte process for separating lignin from plant biomass produces a
class of lignin derivatives called lignosulfonates by attacking the biomass
with an aqueous solution of sulfur dioxide and calcium, magnesium, ammo-
nium or sodium base. Lignosulfonates contain approximately 6.5 weight
percent sulfur present as ionic sulfonate groups on the alkyl chains of the
lignin but are also heavily contaminated with sugars, sugar acids, terpenes,
lignans, and salts. These materials are very water-soluble and commonly have
molecular weights between 10,000 and 40,000. However, molecular weights
up to 150,000 have been obtained for high mass, isolated fractions.
[24]
Less
than 1 teragram (10
12
g) of lignosulfonate are produced in the United States
each year and production from sulte pulping operations is declining year to
year. Environmental restrictions are putting the sulte pulp mills that pro-
duce lignosulfonates out of business. Lignosulfonates for modication or use
MODIFICATION OF LIGNIN 241
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
are being imported into the United States and pulp mills based on anthra-
quinone digestion are beginning to produce lignosulfonates for sale.
A nal extraction process that could become a commercial source of
low molecular weight lignins is The Repap Organosolv Process for removing
lignin in hot ethanol.
[25]
A pilot plant to manufacture 10 megagrams of lignin
a day by this method has been operating in Nova Scotia
[26]
since 1992.
However, the plant, which produced lignin with no sulfur content, an
increased frequency of alkyl groups, and molecular weights of the order of
600 and 1,000,
[27]
has been scheduled for closure in 1997. This leaves only
alkali lignin, sulte lignin, and lignin byproduct of ethanol from biomass
operations available for bulk modication. All other lignins are produced
by processes run on laboratory scale only.
Milled wood lignin (MWL) is produced by grinding wood in a rotary or
vibratory ball mill. Lignin can be extracted from the resulting powder using
solvents such as methylbenzene or 1,4-dioxacyclohexane.
[28]
Milling only
releases 60 weight percent or less of the lignin in wood, disrupts the morphol-
ogy of lignin in wood, and may cause the formation of some functional
groups on the produced lignin.
[29]
Despite these limitations, milling appears
to be an eective way of recovering lignin from plants with only slight
alteration. Enzymes which hydrolyze polysaccharides can be used to digest
plant bers and release lignin. After digestion, the lignin is solubilized in
ethanol.
[30]
Extensive analytical studies support the idea that enzymatically
produced lignin has undergone no major modication in removal from plant
material.
[3135]
Acid hydrolysis of the polysaccharide portion of wood releases lignin
but also causes major condensation reactions
[36]
that remove many ether
bonds in the lignin and replace them with carboncarbon bonds. These
reactions can be minimized by using 41 weight percent hydrochloric acid in
place of other mineral acids but some condensation reactions still occur.
[37]
This is not an eective method by which to obtain unaltered lignin. On the
other hand, lignin can be solvent extracted from wood at temperatures of
175

C using solvent mixtures such as 50/50 (by volume) water/1,4-dioxacy-


clohexane.
[38]
Changes in lignin under these conditions appear to be minor.
C. Uses of Extracted Lignin
Outside of the plant, lignin is useful as a component in the diet of rumi-
nant mammals; a soil property improver in the process of natural decay; and a
source of peat, lignite, and coal. As a commodity forest product, however, it
has a long history as a waste product for which functional uses are sought. This
means that when a woody plant is rendered for its chemical content, about 25
percent of the dry weight of the plant has little or no economic value. For this
reason, the most common use of lignin from pulping operations or ethanol
242 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
from biomass processes is as a fuel. The lignin produced is burned for its
26.5 kJ/g of energy, 40 percent of the solar energy
[4]
stored by the plant.
II. MODIFICATION OF LIGNIN
Other uses for lignin can be broken into three large groups, two of which
require chemical modicationof the biopolymer. These groups are, (1) breaking
lignin down into component aromatics or repeat units, (2) using the biopoly-
mer as extracted from the wood, or (3) adding to the lignin biopolymer,
treating it as a starting material to be built upon to make useful materials.
A. Decomposition of Lignin
Pyrolysis
The decomposition of lignin into aromatic repeat units is a long prac-
ticed art which reached its zenith around 1800 A.D. Production of chemicals
by wood pyrolysis was extensively practiced until, between 1750 and 1850
A.D., coal slowly displaced wood as the major chemical source available to
man. Wood is usually pyrolyzed at 260 to 410

C and lignin at 300 to 440

C to
produce 50 weight percent charcoal, 10 to 15 percent tar, and lesser amounts
of 2-propanone, ethanoic acid, and methanol.
[3941]
The tar is often called
wood creosote, and is a complex mixture of substituted phenols and aro-
matics. It contains phenol, 2- and 4-methylphenol, 2,4-dimethylphenol,
2-methoxyphenol, 4-methyl-2-methoxyphenol, and 4-ethyl-2-methoxyphe-
nol.
[42]
Modications of pyrolysis have been used to convert up to 23
weight percent of starting lignin to ethyne, HCCH, by rapid heating
of the lignin to 1,200

C and rapid quenching of the produced gas.


[43,44]
Ethene, HCCH, is produced in lesser amounts but this synthesis has
signicant potential to produce the largest commodity, organic chemical
made in the world today, ethene, for use in polymerization and synthesis.
This process is not commercialized as of 1997.
Lower temperature pyrolysis has led to lignin-based surfactants. Agroup
at Texaco, Inc., has shown that after retorting the lignin, the phenols can be
ethoxylated to formnonionic surfactants that are both inexpensive and highly
useful in industrial processes such as oil recovery.
[4547]
Alternatively, pyrolysis
in a reducing atmosphere of hydrogen can be used to make cresylic acid in
yields of 35
[48]
to 40
[49]
weight percent of the lignin charged to the reactor.
Cresylic acid is a mixture of alkyl phenols that boils between 180 and 240

C.
Freudenberg and Adam
[50]
were able to convert 35 weight percent of their
starting lignin to a mixture of 10, low concentration, identied phenols and
a large fraction of higher molecular weight, unidentied phenolic products
MODIFICATION OF LIGNIN 243
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
using hydrogen and catalysts to promote the decomposition. Yields have since
been improved
[51]
on reactions on sulte lignin, kraft lignin, soda lignin, and
especially from organosolv lignin. The problem of using such a mixture for
further synthesis of commercial products is the chemical complexity of the
mixture. The majority of all chemicals are used for polymer synthesis and
must be 98%pure. The purication costs for such complex phenol mixtures
will limit their application until catalysts are found that produce 75%,
single-compound products directly from the decomposition reaction.
Reduction with Hydrogen
Hydrogenation over Raney nickel is the most common laboratory
reduction of the structure of lignin into monomeric units. The best conditions
for high yields of aromatic products are 160 to 170

C in 50/50 by volume
aqueous: 1,4-dioxacyclohexane for four to ve hours. This reaction was used
to provide strong support for the C
9
structure of lignin.
[52]
Longer times or
higher temperatures will hydrogenate the aromatic rings of the reaction pro-
duct mixture and/or remove higher levels of methoxyl groups.
Hydrogenation has been used to convert lignin into a liquid that might
be an additive or component of fuels. This or the production of alkyl phenols
is the goal of all eorts to create a commercial process to hydrogenate lignin
into useful products. Two research groups, Inventa A. G. fur Forschung and
Patent Verwertung
[53,54]
and Noguchi Institute of Japan
[55]
have developed
methods to crack lignin to alkanes and phenols. Of these two eorts, the
work done in Japan was carried closest to commercialization by continued
research by Crown Zellerbach Corporation of the United States.
[56]
The
Noguchi process uses a catalyst suspended in phenol or heavy oil under
10.0 to 20.0 MPa of hydrogen pressure at 370 to 430

C to convert 21
weight percent of the original lignin to monophenols. Subsequent work at
Crown Zellerbach increased this yield to 38 weight percent. The product is,
however, a cresylic acid and the preparation of any pure compound by this
process is still highly uneconomical because of costs of separation. Crude
phenolic mixtures produced this way are more economical
[57]
and will prob-
ably be commercialized rst.
Oxidation
The oxidation of lignin
[5860]
in oxygencalcium oxidesodium carbo-
nate, nitric acid, or chrome oxide/ethanoic acid is known to produce 4-
hydroxy-3-methoxybenzaldehyde (vanillin, [121-33-5], equation 1) in amounts
up to 9 weight percent of the lignin. Optimization of the chrome oxide/etha-
noic acid method is claimed
[61]
to produce 4-hydroxy-3-methoxybenzaldehyde
244 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
in 33 weight percent yield. This is one of the few commercial utilizations of
degraded lignin. This food additive and scent is currently produced from
lignosulfonates by Borregaard in Norway and Sanyo in Japan. Organosulfur
chemicals are also produced by retorting lignosulfonates. Dimethylsulde and
methylthiol are produced by this reaction
[62]
with the dimethylsulde being
further oxidized to dimethylsulfoxide (DMSO) for use as a solvent.
This decomposition technology for lignin is emphasized in the literature
as a major means of lignin utilization but it makes little thermodynamic or
economic sense. With two-fths of the plants absorbed energy being used to
make the one quarter of its dry mass that is lignin, lignin represents a large
investment of biochemical eort by the plant. Reducing this macromolecule to
CO
2
or aromatic fragments destroys much of that investment. Keeping the
molecule as extracted from the wood or adding to it are thermodynamically
preferred approaches to lignin utilization but these approaches face signi-
cant, practical barriers. Lignin is a deep brown, uy powder which can be
thermoformed into hard, brittle solids when heated above its glass transition
temperature. This transition from a brittle, amorphous solid to a ductile ther-
moplastic occurs when lignin is heated above 90

C when it contains 13 weight


percent moisture or up to 195

C when it contains 0 weight percent moisture.


Lignin thus changes its properties sharply when relative humidity changes and,
once thermoformed, is a brittle glass at common application temperatures, 20
to 25

C. Further, the deep brown color is a product of free radicals in the lignin
which, if bleached away, will slowly reform by thermal and photo-absorption
mechanisms. These radicals will then react with atmospheric oxygen.
[63]
This
behavior can be a major drawback for applications of lignin to consumer
products. Added to these diculties are the variations in lignin produced by
dierent sources and extraction processes and the chemical complexity already
described. Despite these diculties, the enormous amount of lignin available
at low cost has driven numerous eorts to utilize it.
B. Using Lignin as Extracted from the Plant
Using lignin in the formobtained when it is extracted fromthe plant does
not mean that the lignin exists in the application just the way it does when
withdrawn from the plant. It means that the lignin enters the application
process as a reagent and is often reacted with other components of the
MODIFICATION OF LIGNIN 245
(1)
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
formulation as product is produced. This is denitely the case in the largest
current application for unaltered lignin, its use as a replacement for phenol in
phenolmethanal (formaldehyde) adhesives.
1. Lignin in PhenolMethanal Adhesives
Phenolmethanal adhesives are currently used in about one tenth of all
plywood and particle board. The binding technology represented by these
Bakelite resins would be more widely applied if the reagents, particularly
the phenol, were cheaper. Using lignin in place of phenol will sharply reduce
the cost of the binder.
[64]
Unfortunately, lignin is not structurally equivalent
to phenol. Phenol has 5 hydrogen sites on the aromatic ring and no non-
hydrogen, ortho or para substituents around the hydroxyl group. Kraft pine
lignin has only 72 phenolic hydroxyl groups per 100 C
9
repeat units and 48
aliphatic hydroxyl groups per 100 C
9
repeat units.
[65]
For virtually all lignin
phenolic hydroxyl groups, the aromatic ring is para substituted by the propyl
chain of the 1-propylphen-4-ol (coumaryl) structural unit. In softwoods, the
hydroxyl group is often next to a methoxyl group in the number 3 position on
the ring while in hardwoods, it is completely ortho substituted by methoxyl
groups. This leaves only the meta position open for reactions on the aromatic
ring of a lignin phenol. The implications of this structure on lignin reactivity
in phenol/methanal crosslinking polymerizations can be seen from the
mechanism of the phenolmethanal reaction, shown in Figure 2.
In crosslinking with methanal, an aromatic hydroxyl group ionizes to
form ortho [2,6] and para [4] anionic sites through which to react with a
positively charged, methylene group. Lignin has most sites ortho and para to
its aromatic hydroxyl groups blocked by organic, functional groups. This is
the reason why lignin reacts slower with methanal than does phenol and why
lignin can only be used to replace between 40 and 70 weight percent of the
phenol in an adhesive formulation. Lignin simply has too few, highly reactive
sites to create a high density of crosslinks without at least 30 weight percent
phenol being present. The rates of reactions determined by Dr. Douglas
Gardner are compared between hardwood, steam-exploded lignin; softwood,
kraft lignin; and phenol
[66]
in Table 5-A. The rate of the hardwood lignin/
methanal reaction is, as would be expected from the dimethoxyl substitution
on the ring, only 46 percent as fast as phenol at 30

Cand only 12 percent as fast


at 60

C. Softwood lignin has, under the same reaction conditions, a rate that is
68 percent as fast as phenol at 30

C and 14 percent as fast at 60

C. The open
ortho positions on softwood lignin obviously allows the softwood lignin to
react more readily with methanal and should lead to more extensive cross-
linking of the softwood lignin as compared to the hardwood lignin. This is
conrmed by the data of Table 5-B. Here the number of methanal groups
added to each C
9
repeat unit of the two lignins is determined by three dierent
246 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
MODIFICATION OF LIGNIN 247
Figure 2. Mechanism of the phenolmethanal reaction.
Table 5. Kinetic Parameters for the Phenol/Methanal or Lignin/Methanal Reactions Rate
Constant
k(10
2
M
1
min
1
)
Temperature,

C
Pre-exponential
Factor
A (min
1
)
Activation
Energy
E
a
(kJ/mole) A: Component 30 40 50 60
Phenol 2.17 6.65 24.9 79.4 5.8 10
15
101.3
Kraft lignin 1.44 1.83 5.50 11.3 3.25 10
8
60.7
Steam-exploded lignin 0.98 1.60 4.37 9.44 1.32 10
9
64.9
B: Degree of Methanal (HCHO) Substitution per C
9
Unit of Lignin by Various Methods
a
Degree of Methanal Substitution
Method Kraft lignin Steam-exploded lignin
HCHO uptake 0.39 0.25
1
H-NMR
b
0.38 (0.35) 0.18 (0.150.20)
13
C-NMR 0.42 0.27
a
Data from Reference 49. Formula for the rate of reaction, k, is KA * e
(E
a
/RT)
.
b
Values in parenthesis from Reference 66.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
methods. The data show that hardwood lignin only reacts with 0.23 methanal
units per C
9
while softwood lignin reacts with 0.40 methanal units.
With adjustments in composition to compensate for the chemical fea-
tures of each aromatic hydroxyl source, a wood binder formulated with
hardwood lignin, softwood lignin, or phenol
[67]
will be deemed highly eec-
tive if it can be: (1) formulated at lower cost, (2) applied with conventional
equipment, (3) reacted under the same process conditions, and (4) an adhe-
sive that is so strong that wood parts formed with it fail in the wood phase
most of the time and not in the adhesive phase.
Adams and Schoenherr
[68]
achieved most of these bench marks by for-
mulating an adhesive consisting of a 40 weight percent solids solution of kraft
lignin in phenol/methanal/sodium hydroxide. This uid had a viscosity of
10 Pas and thus was a very thick and energy-consuming adhesive to spread.
However, when this binder was used in the manufacture of three ply panels of
Douglas r, destructive testing of the plywood showed failure in the wood
phase 92 percent of the time. A more easily applied adhesive can be prepared
by blending 37 weight percent lignin in phenol/methanal/sodiumhydroxide
[69]
and only partially crosslinking the mixture. This blend has a viscosity of
0.46 Pas but sets into an adhesive layer under 1.2 MPa pressure for 6 minutes
at 140

Cthat breaks in the wood phase 94 percent of the time. These data show
that, despite its chemical deciencies, lignin is a functional replacement for
much of the phenol in Bakelite adhesives. Appropriately blended, lignin-
containing adhesives will, under common treatment conditions for binding
plywood or particle board, set into an adhesive that is stronger than the
wood
[70]
and therefore, capable of producing bonds that will be the last part
of the structure to fail. As of 1991, lignin constitutes 17 percent of the resin
solids in phenol-methanal adhesives used to make exterior-grade plywood.
[71]
The lignin used is generally kraft lignin. This technology is providing a small
but stable market for the lignin fraction of wood. Growth areas of adhesive
bound, wood composites; oriented strandboard; oriented waferboard;
mediumdensity berboard; and laminated veneer lumber will provide a grow-
ing market for lignin as a partial phenol replacement. Organosolv lignin is also
being used as a resin extender for high performance markets of phenol/
methanal resins. Organosolv lignins are used in brake pad and foundry
binder resin formulations of phenol and methanal.
2. Lignin Photostabilizers
DePaoli and Furlan have studied the use of lignin from sugar cane
bagasse as a photostabilizer for butadiene rubber.
[73]
The logic for this appli-
cation is that the phenol-containing repeat units of lignin have structures
proximate to compounds currently known to act as photostabilizers in
rubber. Hindered alkyl phenols with long chains para to the hydroxyl
248 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
group, structure 1 in Figure 3, are known to inhibit photo-induced bond
cleavage in rubbers by forming stable phenoxyl radicals.
[74]
This stable, hin-
dered radical prohibits the formation of a peroxide radical on the rubber
backbone, thereby preserving the structural integrity of the elastomer.
Bagasse lignin contains approximately 2 weight percent of structures
2 and 3 with the frequency ratio between them being 4 of structure 2 to 1 of
structure 3. These structures are not only similar to those of common photo-
stabilizers, these repeat units appear in a lignin chain. Polymer-bound,
hindered phenols are more eective than free, molecular phenols because
the polymer chain restricts migration and dimerization of the formed radi-
cals.
[75]
Bagasse lignin was tested as a mixture of 90 weight percent lignin and
10 weight percent N
0
,N-bis(1-ethyl-3-methylpentyl)-p-phenylenediamine in
butadiene rubber. Diamines are commonly used in conjunction with hindered
phenols to inhibit photodegradation in rubber. Rubber samples containing
the lignin blend and commercial stabilizers were irradiated at 350 20 nm in
air and rates of photodegradation were measured. The data showed that 0.37
weight percent diamine could be replaced by 2.25 weight percent lignin with-
out aecting photostability of the blend. The lignin stabilized the rubber by
both capturing radicals and absorbing the ultraviolet light directly. The eect
on physical properties of compounding butadiene rubber with over 2 weight
percent lignin was not investigated. While these data are positive, they fail to
verify that the photostabilized rubber possesses all of the application proper-
ties that the rubber must have to be used as a commercial product. Lignin has
long been known to be an excellent reinforcing agent for rubber if the low
MODIFICATION OF LIGNIN 249
Figure 3. Base structure of photoinhibitor (1) and common bagasse lignin repeat units (2,3).
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
molecular weight lignin and non-lignin constituents of industrial byproduct
lignin are removed.
[76]
If contaminants are not removed from industrial
lignin, they promote clustering of the lignin particles,
[77]
lowering of the soft-
ening temperature of the rubber, and lowering of the reinforcing ability of the
lignin. Unfortunately, the bagasse lignin used in the photostability studies
was only 93 weight percent lignin and was not fractionated to remove low
molecular weight portions of the blend. Because of these deciencies, the
photostability data are of limited signicance.
3. Lignin in Electrodes and Lignin Electrodes
Lignosulfonates are used in every lead/acid battery to inhibit crystalliza-
tion of the negative electrode during recharging of the battery.
[78]
Without the
addition of a puried, modied lignosulfonate to the negative terminal of the
battery, the battery would fail to charge after only a few discharge cycles.
[79]
This lignosulfonate is commonly called an expander and is thought to con-
trol the formation of lead sulfate
[80]
on the surface of the sponge lead in the
negative terminal.
Lignin pyrolyzed at 700

C under nitrogen forms a cohesive, conducting


solid which can act as an electrode in a storage battery. This modication to
dehydrogenate and deoxygenate the lignin forms a charcoal with the capacity
to absorb or donate 6 mmoles of electrons per gram.
[81]
Batteries have been
formed from these electrodes and the cells have produced 45 W-hr/kg per
electrode at 70 percent eciency (charge recovered/charge put in). While the
watt-hours/kg rating of these electrodes is about two thirds of the value of a
lead oxide plate, the lignin-based electrodes polarize rapidly and suer a
rapid drop in discharge voltage.
[82]
These two performance properties work
against eective use of the electrodes. Further, the redox capacity of carbon
structures is quite limited when compared to metals so the utility of this
modication has yet to be veried. However, since the internal structure
and composition of these electrodes are unknown and, under current tech-
nology, controlled solely by pyrolysis conditions, there is extensive room for
improvement of these biomass electrodes. The drive to improve these elec-
trodes will be promoted by the fact that carbon is only 18 percent as heavy as
lead, the common electrode for transportation batteries.
4. Construction Binders
A limited amount of lignin is used as the binder for glass wool build-
ing insulation. It is applied to the hot glass as the ammonium salt solid
recovered from the kraft paper production process and allows the glass
bers to bind to one another when the ber pads are spun or formed.
250 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
5. Food Additives
A possible use for lignin in the future is the addition of alkali lignins to
pet and human food as a roughage or ber source.
[8385]
Extensive research
has shown that high dietary ber correlates with low incidents of colon
cancer.
[86,87]
The lignin in biomass is known to be a signicant source of
this cancer protection,
[88]
possibly producing this eect by trapping free radi-
cals while in the colon.
[89,90]
If further work shows no adverse eects from
lignin, it may be added to food to increase long term, public health.
6. Fat Purication
Small amounts of free acid lignin, a kraft lignin with its anionic
groups neutralized with hydrogen ions, are used to purify recycled animal
fat. The yellow grease from food preparation is puried by ltration
through the lignin and the contaminated lignin is sold for use in animal feed.
C. Adding to the Lignin Biopolymer
The reactions used to modify polymeric lignin are:
. Alkylation and dealkylation . Sulfomethylation
. Methylolation . Sulfonation
. Amination . Nitroxide formation
. Carboxylation and acylation . Silylation
. Halogenation and nitration . Phosphorylation
. Hydrogenolysis . Grafting and oxyalkylation
. Oxidation and reduction
A large number of lab scale, lignin modications were performed solely
to determine the structure of lignin. Most of this work was done before 1945
and represents a classic approach to the verication of structure by synthe-
sis of an identiable adduct or derivative of the structure being sought. The
reactions run to verify structure were: formation of ethanoate esters to iden-
tify hydroxyl groups, capping of hydroxyl groups with dimethylsulfate to
form methoxyl groups, demethylating methoxyl groups with hydrogen
iodide to form hydroxyl groups, oxidation with nitric acid to decompose
and/or add to double bonds with halogens, and reacting with copper
[1]
oxide to identify aldehydes.
The investigation of ax lignin by Powell and Whittaker
[91]
is a strong
example of many eorts in this type of search for structure through modica-
tion.
[9297]
While these eorts did identify the C
9
repeat unit of lignin, the
essential goal of this work to identify a repeat unit structure common to all
lignins is now known to be spurious. The work did help to establish that
MODIFICATION OF LIGNIN 251
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
dierent species and plant parts have lignins containing dierent amounts of
functional groups and dierent repeat unit bonding. Laboratory-scale mod-
ications of lignin will be cited repeatedly in the discussion of modication
which follows.
1. Alkylation
The three methods of alkylating lignin are:
a. reaction with diazoalkanes,
b. reaction with alcohols in the presence of a catalyst, usually hydro-
chloric acid,
c. use of alkylsulfates and sodium hydroxide.
The sites of alkylation are the oxygen atoms of the hydroxyl, carbonyl, and
carboxyl groups in lignin. Reactions (a) and (b) are selective with diazoalkanes
(a) reacting under anhydrous conditions to alkylate, chiey, the slightly acidic
hydroxyl groups
[98]
of the phenolic, enolic,
[99102]
and carboxylic
[103107]
units
to form ethers. The diazoalkane reaction with carboxylic acids only occurs in
solvents in which the acid is deprotonated to an enolate anion. The RN
2
reactions are shown in equation 2.
A catalyzed alcohol reaction (b) on lignin takes place at a-hydroxyke-
tones, carbonyl, and carboxyl groups. The reactions give dierent products
and are shown in equation 3. Since alcohol/acid alkylation does not alkylate
the aromatic hydroxyl groups,
[108]
alkyllignins from these reactions have
sharply dierent solubility and physical properties than those from diazoalk-
252 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
ane alkylation. They are, for example, soluble in dilute, aqueous sodium
hydroxide and are generally more hydrophilic than diazoalkane-treated lignins.
The use of dialkylsulfate, R
2
SO
4
, and base to alkylate lignin produces
a product in which the aromatic and primary and secondary, aliphatic
alcohols
[109,110]
are alkylated. Some tertiary alcohol groups may escape
alkylation.
[111]
These products are extensively hydrophobic since almost
all of the acidic alcohol groups are now capped as ethers and all carboxylic
acid groups have been converted to esters. Unless other ionic functional
groups are present, these products dissolve only in nonpolar or organic
solvents and have the physical properties of a thermoplastic.
Dealkylation is a common byproduct of a number of reactions on lignin
and will be treated under those substitution reactions.
2. Methylolation
The addition of a methylol group, CH
2
OH, is a result of the addition
of methanal, H
2
CO, to the aromatic rings of the polymer. This reaction
has been investigated as a stepping stone to the use of lignin in place of
phenol
[112,113]
in phenol/methanal resins (See Chapter 12). As detailed in
Section II.B.1, the reaction of methanal on lignin is hindered by the presence
of ortho [#3-position] methoxyl groups and para [#1-position] alkyl groups
on the only aromatic rings that will engage in this addition reaction, those
MODIFICATION OF LIGNIN 253
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
with a free phenolic group. These phenomena are illustrated in Figure 2, the
mechanism of the methanolylation.
The level of methanolyl group addition depends on reaction pH and the
source and extraction process used to recover the lignin. Morton and Adler
[114]
show that under alkaline (pH>7) conditions, milled wood spruce lignin takes
up 0.15 CH
2
OH groups per methoxyl group while kraft pine lignin takes up
0.5 CH
2
OHgroups per methoxyl group. Without free phenol present, Allan
and Halabisky
[115]
claim that this reaction does not lead to the increase of the
limiting viscosity number, [], of the lignin during reactions to extensively
methanolate lignosulfonates. The failure to increase [] would mean that the
molecular weight of the lignin sample is stable and that intermolecular methy-
lene bridges, shown in equation 4, fail to form. In equation 4, R
3
and R
4
are the
linkages to the rest of two distinct, lignin molecules. However, both research
and process work at the Georgia Pacic Corporation
[116]
proves that ligno-
sulfonates can be crosslinked into a solid by methanal under both acid and
alkaline pH. It would appear likely, therefore, that the studies by Allan and
Halabisky
[115]
are wrong and methylene bridges between lignosulfonate mole-
cules can be made in alkaline solutions by methanal.
3. Amination
Amination, the creation of a NR
0
2
group on lignin where R
0
H or an
organic unit, is a reaction used to identify structures and create emulsion sta-
bilizers. The reaction is used to prove the presence of aromatic nitro groups,
NO
2
, in lignins treated with nitric acid.
[117]
The nitrolignin has been treated
through reduction with Raney nickel,
[118]
zinc and acetic acid,
[119]
and sodium
amalgam
[120]
to forman amine; the amine has been converted to a diazo group,
N N; and the diazo group has been reacted with substituted phenols to
formdyes. There is no commercial utilization of the dyes that can be formed by
this sequence of reactions but the formation of diazo dyes does verify the
presence of nitro groups in the material treated to form the amine.
Reactions have been run to make amine-group-containing lignin
[121123]
and free, amine groups which can accept a proton have been quantied
[118]
in
the lignin products. The use of amination to form commercial products of
lignin requires that beta amino ketones be formed on the aliphatic chains in
the polyalkylaromatic. The acid catalyzed addition of methanal and a primary
or secondary amine (Mannach reaction
[124]
) is conducted on alkali lignin in
aqueous suspension. The reaction is shown in equation 5. In a typical synthesis
reaction
[125,126]
with all percents by weight, 19 percent alkali lignin, 77 percent
water, 2.3 percent dimethylamine hydrochloride, and 1.4 percent methanal are
slurried at a pH of 2 to 3 and heated to 80 to 85

C for 4 hours. The reaction


mixture is diluted with an equal volume of water and pH is reduced below 1
before the mixture is boiled briey. The product is a water soluble, beta keto
amine derivative of lignin.
254 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
The resulting lignin with beta keto amine groups in it is used in con-
centrations of 0.2 to 2.0 weight percent to form thermally stable, aqueous
emulsions of 55 to 65 percent asphalt for use in road repair. The pH of the
emulsion is between 9.5 and 10.5. It can be used as a wetting binder with
either siliceous or limestone aggregates to form bituminous slurry seal coat-
ing in the stabilization treatment of paving bases or can be used alone in the
manufacture of oor mastics.
4. Carboxylation and Acylation
Carboxyl and acyl groups; a and b, respectively in equation 6, are
common components of lignin, as shown by the data of Table 4. The fre-
quency of both groups is increased by oxidation of lignin and stronger
hydrophilic and polyelectrolyte behavior can be produced in lignin by
these reactions. Typical oxidations that produce sharp increases in the diva-
lent oxygen bond content (oxo units) of lignin are treatments with chlorine,
MODIFICATION OF LIGNIN 255
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
nitric acid, and peroxides. These treatments are usually associated with
extensive dealkylation. Shorygina treated hydrolysis and hydrochloric acid
lignin with gaseous chlorine in 10 percent, aqueous hydrochloric acid to
introduce approximately two chlorines per C
9
repeat unit in the lignin.
However, the carboxyl content of the lignin increased by a factor of between
10 and 20 and 66 percent of all methoxyl groups
[127]
were removed from the
lignin. Similarly, after treating r hydrochloric acid lignin with nitric acid in
tetrachloromethane, Shorygina found a 10 to 20 fold increase in the carbonyl
content of the lignin
[128]
and numerous authors have documented a drop in
methoxyl content of nitric acid treated lignin.
[119,120,127]
By use of a peroxide under alkaline conditions, Bailey and Dence
[129]
produced a sharp increase in the carbonyl content of lignin but also induced a
16 to 17 percent loss of methoxyl groups as methanol. Further, production of
ethanedioic, 1,3-propanedioic, 1,4-butenedioic, and 2-methoxy-1,4-butane-
dioic acids veried extensive degradation of the lignin. This degradation
occurred at phenolic units, ether bonds, carbonyl units, and phenylprop-2-
enal units in the lignin, which is structure c in equation 6. In aqueous, alka-
line peroxides,
[130]
hydroperoxy anion, HOO

, attacks phenylprop-2-enal to
form phenylmethanal and a 2-hydroxyethanal fragment. Such a reaction on
phenylprop-2-enal units in lignin would produce a signicant increase in
carbonyl group content.
As of 1997, only carboxyl and acyl group formation on lignin by chlorine
or alkaline peroxide has become a commercial process
[131]
and that only as a
side eect of chlorine and peroxide bleaching of pulp for paper and packaging
applications. The use of chlorine in this process is declining because it also
produces chlorinated, organic byproducts. The critical reaction that these
bleaching operations produce is not the formation of oxo groups, >CO,
but the removal of the approximately 10
18
free radicals per gram of lignin
[132]
in mechanically or chemically rendered, wood pulp. Terminating these free
radicals removes the intense, broad absorbance band centered at 284 nm that
gives a brown color to the pulp. The removal of these free radical absorbers
thereby brightens (whitens) the pulp. There is some research underway to
create oxo groups in lignin by reaction with ozone, O
3
. Ozonation to induce
carbonyl or acyl groups is not a commercial process in 1997, however.
5. Halogenation and Nitration
One of the simplest addition reactions to lignin is the addition of a halide
to the alkylaromatic backbone of lignin. The reaction is given in Figure 4. The
reaction is run by bubbling chlorine into spent, aqueous pulping liquor and
following that addition with additions of bromine and chlorine.
[133]
The weight
percent halide in the product is raised to between 20 and 40 percent. Since the
halogenated alkylaromatic is hydrophobic, it precipitates. The molecular
256 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
weight of several halolignins has been determined by collegative methods
(cryoscopically
[134]
; ebulliometrically
[127]
) and found to be between 1,000
and 6,000. The halogination reaction is accompanied by extensive dealkylation
and breaking of ether links in the lignin.
[127,135]
This second reaction would
sharply lower the molecular weight of the treated lignin.
Previously, the halogenated lignin has been recovered as a re retardant
for use in building materials and consumer goods. By 1997 standards, how-
ever, this halogenated organic presents very signicant environmental prob-
lems for virtually any application and it is very improbable that this
chemistry will continue to be used today or in the future. Indeed, several
lignin producers specically avoid deliberate or unintended halogenation of
any wood product just to avoid environmental problems that these com-
pounds could produce.
[136]
Nitration is another fairly simple reaction that is typically run in non-
aqueous solvents to nitrate lignin in wood meal and run in suspension or
solution for isolated lignins.
[137,138]
Typical nitrating agents are nitric acid in
concentrated ethanoic acid
[139141]
and nitric acid with sulfuric acid or oleum
(fuming sulfuric acid). The resulting amorphous, yellow to brown powders
have molecular weights of 600 to 2,000 and are extensively dealkylated and
degraded.
[142145]
Nitrogen content of the product may be as high as 6.7 weight
percent.
[146]
The nitrogen-containing structures that are produced in lignin are
shown in equation 7. The nitro groups, 7a., are usually attached to aromatic
rings and often constitute 70 to 80 percent of the nitrogen.
[139]
The nitroso
groups, 7b., are uncommon and seldom represent more than 2 percent of the
MODIFICATION OF LIGNIN 257
Figure 4. Formation of halolignin.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
organically bound nitrogen.
[147]
In the presence of excess nitric acid
[148]
or in
acidic, anhydrous, nitrating conditions,
[142]
nitrate ester groups, 7c., can
be added to lignin. There is no commercial application for nitrated lignins
as of 1997.
6. Hydrogenolysis
The hydrogen in lignin is increased by a spectrum of reactions ranging
from catalytic reaction with hydrogen gas, H
2
(g), to reaction with sodium
borohydride, NaBH
4
. These reactions are conducted for one of two reasons, to
determine the structure of lignin or to convert lignin to a liquid. Hydrogen
addition to determine structure is a research process and has been conducted
on a lab scale for 70 years. Hydrogen addition to liquefy lignin is an eort to
convert lignin to products that could be added to fuels or be a petroleum
replacement. It is covered under the topic Reduction with Hydrogen in
Section II.A. Here, a representative group of hydrogenation reactions which
provide structural insight and specic alterations in lignin will be described.
For a more detailed analysis of hydrogen addition to lignin, see Reference [150].
Typical hydrogen reductions for laboratory structure evaluation would
be reaction of lignin with lithium aluminum hydride; sodium borohydride; or
hydrogen over a catalyst, such as Raney nickel in an aqueous-organic solvent
mixture. Lithium aluminum hydride is a strong but selective hydrogenating
agent and will reduce aldehydes, ketones, acids, and esters to alcohols with-
out reacting with carboncarbon double bonds in the lignin. It is reacted with
lignin in anhydrous ether at room to elevated temperature. Its reaction (8a.)
and that of the other lab hydrogenating agents are shown in equation 8.
Sodium borohydride is a weaker hydrogenating agent and will reduce alde-
hydes and ketones to alcohols without reacting with acids, esters, and
carboncarbon double bonds in the lignin. It is reacted with lignin in alkaline
solution at room to elevated temperature. Since both ring and aliphatic
carboncarbon double bonds and ether bonds are spared reaction with
these two agents, the three dimensional structure of lignin is not eected
by the reduction and important structural features, such as the frequency
of conjugated, carbonyl bondings, can be measured. Adler and Marton
[151]
used this feature of the chemistry of these hydrogenating agents to calculate
the frequency of carbonyl groups adjacent to aromatic and ethene groups.
This data is shown in Table 6.
7. Oxidation and Reduction
Oxidation, the loss of electrons, and reduction, the gain of electrons, are
labels for two major classes of reactions. The reactions described here that
258 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
are oxidations are alkylation, sulfomethylation, methylolation, sulfonation,
amination, nitroxide formation, carboxylation, acylation, halogenation,
nitration, phosphorylation, oxyalkylation, and, generally, grafting. The reac-
tions that are reductions are dealkylation, silylation, and hydrogenolysis.
MODIFICATION OF LIGNIN 259
Table 6. Estimated Conjugated Carbonyl Groups in
Spruce, Milled Wood Lignin
Functional Group Structure
Percent of all C
9
with This Structure
<1
3
<1
5 to 6
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
8. Sulfomethylation and Sulfonation
Sulfomethylation and sulfonation are reactions to add the methylene
sulfonate, CH
2
SO

3
, and sulfonate, SO

3
, group, respectively, to lignin.
The addition of methylene sulfonate groups to lignin has been proposed as a
means of improving the tanning capacity of lignin.
[152]
The reaction is run
with equal moles of methanal, alkali metal sulte salt, and reactable phenolic
repeat units in water under neutral to basic pH and 100

C temperature.
Repeat units of lignin which can react must have either a hydroxyl group
on the aromatic ring or a carbonyl group alpha to a double bond or the
aromatic ring. These conjugated carbonyl structures are shown as structures
a and d of Table 6. These repeat units are common in lignin and produce
polyelectrolyte products with some repeat units in the lignin having the
structures shown in equation 9a for aromatic hydroxyl groups and 9b for
conjugated carbonyl groups, respectively. Sulfomethylation, as opposed to
sulfonation, is only used when an increased content of sulfonate groups is
needed in the lignin product. It is applied to kraft lignins by the Westvaco
Corporation to form dye dispersants that are marketed under the Reax trade
name. The presence of base in the reaction will lower the yield of sulfonate
groups and increase the number of hydroxymethyl groups
[153,154]
added to
the lignin product.
Lignin sulfonation is the most studied reaction in lignin chemistry since
it was the earliest and cheapest way to make commodity, acidic cellulose for
paper. The sulte process, described in Section I.B, was patented
[155]
in 1866.
Research on the reaction began in the late 1880s and produced the conclu-
sion, by 1892, that the process incorporated sulfur into the lignin and that
sulfur was present as sulfonic acid groups.
[156158]
Holmberg showed in 1935
that lignosulfonates
[159,160]
were made by reactions at hydroxyl groups alpha
260 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
to aromatic rings, as shown in equation 10a, with these reactions being
promoted by free phenolic hydroxyl groups, R
4
H. Sulfonation occurs
rapidly at ether linkages alpha to a phenolic ring, equation 10b; hydroxyl
groups gamma to a phenyl carbonyl pair, equation 10c; and slowly at alpha-
beta diethers o the aromatic ring, equation 10d. The lignosulfonates are the
most extensively used lignin product with a teragram (10
12
g 10
9
kg) of this
material sold in the United States every year. Industries which use lignosul-
fonates are well drilling; cement manufacture, formulation, and pouring;
ceramics manufacturing; and construction materials.
Both lignins and lignosulfonates are used to prepare oil well drilling
muds but lignosulfonates control the majority of the market. The lignosul-
fonate must disperse and hydrate the clay in the mud but not perform these
actions on the formation being drilled. The drilled solid must be maintained
as particles and suspended in the mud until it can be ltered out on the
surface before the mud is recycled into the well.
[161]
This requires that the
mud be a gel when static but a mobile uid when the drill bit is moving. This
means that the lignosulfonate must make the mud thixotropic (shear thin-
ning). The mud must be stable to all formation contaminates and brines and
at temperatures up to 200

C. Lignosulfonates oxidized by chromate or


dichromate ion or mixtures of these ions and iron salts perform all of these
functions cost eectively.
[162]
The metal ions oxidize the lignosulfonate mole-
cules to produce more carbonyl and acyl groups and to provide di- and tri-
valent cations to temporarily crosslink carboxylate anions into molecular
bridges to create transitory molecular weight increases in the polymer in
the static mud. These materials are marketed as chrome or ferrochrome
lignosulfonates.
The biggest industrial user of lignosulfonates or sulfonated lignins is the
cement industry, which uses these compounds as grinding aids, air entraining
agents, concrete additives, and grout hydrating agents to control the setting
and hydration rate of cement. When cement is produced from a kiln, the
clinker must be ground into the ne powder that will be sold.
Lignosulfonate and sulfonated kraft lignin are added to the clinker during
grinding to inhibit reagglomeration of the ground particles. The polyaro-
matic, sulfonated polymer apparently reacts with bonds broken in the grind-
ing to inhibit rebonding between particles. As little as 200 ppm by weight can
prohibit reagglomeration of the cement.
[163,164]
When cement is poured, it must contain a certain amount of air to
prevent spalling (large particle formation) and cracking. The air bubbles
act as internal interfaces to inhibit the propagation of cracks.
Lignosulfonates are used as air entraining agents to form these bubbles
and increase
[165]
the strength of nished concrete. Adding as little as 0.3
weight percent of a mixture of lignosulfonate and cement to concrete and
cement as it is being mixed or placed allows a reduction in water use in the
cement of up to 20 percent. The cement has a higher compressive strength
MODIFICATION OF LIGNIN 261
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
and durability than non-lignosulfonated cement for the rst several years
after pouring,
[166]
so the additive induces quick strength and durability in
the solid.
Sulfonated lignins are dispersants for the preparation of pastes and
slurries of solids in water. Since the typical lignosulfonate contains only
one sulfonate group per two to three C
9
repeat units, the molecule is a
surfactant structure with a 15 to 25 carbon atom, hydrophobic section.
This alkylaromatic section of the molecule will plate onto solid surfaces in
a slurry and by so doing, will impart a negative charge to the particles. The
like-charged particles of the slurry or paste now become electrostatically
repulsive to one another and the suspension is not only more stable, it has
262 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
signicantly lower viscosity.
[167]
Lignosulfonates for this application have
molecular weights of between 10,000 to 40,000
[168]
and are most commonly
used in the ceramics and clay pottery industry at concentrations of 0.05 to 0.4
weight percent of the clay solids used in a formulation. The lowered viscosity
of the clay slip allows reduced levels of water to be used in the clay formula-
tion, reduces energy costs in ring, and produces objects with increased green
and red strength.
[169]
In the formation of building brick, the dispersing
properties of lignosulfonates are applied in the same way they are in other
ceramic applications and with the same eects. However, the lignosulfonates
also lubricate the plug mill and reduce energy costs for brick molding.
Application levels for lignosulfonates in brick are between 0.5 and 2 weight
percent of the unred clay formulation.
[170]
Formation of wallboard for
MODIFICATION OF LIGNIN 263
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
construction uses 0.05 to 0.16 weight percent lignosulfonate to disperse the
gypsum or its partial hydrates in the slurry from which the panels are
made.
[167]
The largest single use of lignosulfonates is their application to dirt roads
for dust control and road stabilization.
[171]
The lignosulfonate liquor con-
tains, on average, 10 weight percent hexose sugar, 10 percent pentose sugar,
and 10 percent non-cellulosic carbohydrate. It is this fraction of the byprod-
uct lignosulfonate from paper manufacture that acts as a binder in the road
surface, reducing dust and increasing the road capacity to withstand heavy
trac, resist erosion, and minimize frost heave.
[172]
Agricultural slurries such as pesticides, herbicides, and fertilizers are
dispersed with 2 to 5 weight percent lignosulfonates.
[173,174]
Similarly,
carbon black slurries in latex rubber are dispersed to produce a more uni-
form spread of the carbon particles in the precipitated rubber used in
tires.
[164]
Sulfonated lignin is used in the mining industry to rene ore by ore
otation. The lignosulfonates depress the entrainment of minerals such as
calcite, barite, and talc and produce a higher metal content extract from the
ore.
[167]
Lignosulfonates were added to water in cooling towers or heating/air
conditioning systems and were also used to treat boiler water. The low cost of
lignosulfonates plus their ability to sequester calcium or magnesium ion
[175]
or disperse heat coagulable particles
[176]
produced a signicant market for
these anionic polymers that peaked in market share and volume in the 1960s.
Applications levels were 0.1 weight percent or less. Lignosulfonates have
been almost completely replaced in these applications by copolymers based
on 2-propenoic acid or amide. If lignosulfonates are to be used again in
cooling towers or boilers, the materials must be cheaply modied to powerful
scale, corrosion, and deposition inhibitors.
With further modication, lignosulfonates are the principle primary
dispersant, extenders, and grinding aids for the dye industry. Lignosulfonates
reacted with 1-methanolylphenol (second reagent, equation 10e) have
[177]
increased thermal stability, good grinding aid performance, high dispersion
eciency, low ber staining behavior, and a minimal tendency to chemically
reduce the azo group in azo dyes. For these reasons and their low cost,
methanolylphenol-lignosulfonate esters are extensively used in the dye indus-
try. The modied lignosulfonate is made by reacting 0.5 mmoles of 1-metha-
nolylphenol per gram of lignosulfonate in a 50 weight percent, aqueous
solution at a pH of 10 and a temperature of 100

C for 5 hours to form an


esteried lignosulfonate.
[72]
The reaction is illustrated in equation 10e.
During this reaction, aldehyde groups in the lignosulfonate may be oxidized
to carboxylic acid groups, a transformation illustrated by having structure
6b, with RH, of equation 6 become structure 6a. This removes reducing
groups from the lignosulfonate.
264 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
9. Nitroxide Formation, Silylation, and Phosphorylation
These three modications are all research techniques or laboratory
chemistry that have been done on lignin. The methods have no commercial
use as of 1998. Nitroxide addition to lignin is a reaction done solely to elicit
structural information about lignin by introducing free radicals into the
polymer. The functional group added to lignin is shown in equation 11a
where R
1
is often a phenyl group of lignin, ester-linked to the nitroxide, and
the period superscript represents the unpaired electron on the nitroxide. As
the structure implies, nitroxide groups are often added to molecules contain-
ing hydroxyl groups through esterication reactions using the carboxyl group
on the most common reagent used for nitroxide addition, 3-carboxy-2,2,5,5-
tetramethylpyrroline-l-oxyl radical, equation 11b. The addition of 11b. to
three lignin model compounds; 2-methoxyphenol, 4-(1-ethanalyl)-2-methoxy-
phenol, and 4-(1-methanolyl)-2-methoxyphenol; to make structures of type
11a. showed
[178]
that the electron spin resonance spectra of the three, radical-
containing molecules was essentially the same. This removes the capacity to
form these radicals on lignin and deduce structure and bonding in the natural
polymer. The authors try to promote the usefulness of this synthesis and ESR
test because diradicals on the same lignin repeat unit show ESR signals that
are characteristic of the distance between the radicals. Unfortunately,
1,n-benzenediol groups, where n ranges from 2 to 4, are infrequent in lignin
and the frequency of these groups and the distance between the hydroxyl
groups on them does not constitute critical structural information. This
explains why this method has not been applied again to examine lignin
structure. Nitroxide adducts of lignin have been formed
[179]
to examine the
coil behavior of lignin in solution but the study provided minimal technical
insight. While the free radical probes on solvated lignin molecules did show
dierences in dierent solvents, radical concentrations, and solution concen-
trations, the dierences were not readily interpreted in terms of the structure
of the coil and the interactions between the molecules. The dispersion and
solvation of lignin in solvents, particularly aqueous base, is of general impor-
tance in lignin chemistry and the digestion of wood pulp. However, the study
MODIFICATION OF LIGNIN 265
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
of solvated nitroxide radicals performed by Lindberg has not been repeated
and, when used with 1998 technology, this method is apparently incapable of
answering questions about aggregation, coil dispersal, and polymer-solvent
interaction parameter that are of interest in lignin chemistry.
Silylation requires that bonds be formed between lignin and a silicon
atom. This type of reaction has been claimed in the literature
[180]
but the
labels applied and chemistry assumed are invalid. In a patent, Blount claimed
that a reaction between sodium silicate, Na
2
SiO
3
, sulfuric acid, and lignin
produced gray, granular solids labeled lignin silicoformate and lignin silicate.
The reaction is conducted by mixing sodium silicate and sulfuric acid in
approximately stoichiometric quantities to form silicic acid,
[181]
a metastable,
aqueous solution of Si(OH)
4
which will react into colloidal silicates and
polymerize at a rate dependent on pH.
[182]
If lignin were introduced to this
solution, the forming silica network would entrap the lignin macromolecules
and form, depending on lignin concentration, anything from an adsorbed
lignin in a silicate gel to an interpenetrating network of silicate gel and lignin.
If this forming or formed gel were heated to evaporate water, the dehydrating
mixture of a silicate lattice and an alkylaromatic polyalcohol would react to
release water by forming alkoxide bonds to silicon, as shown in equation 12.
In isolation, this alkoxide bond would form rapidly with a phenolic hydroxyl
group but hydrolyze equally rapidly when re-exposed to water. Aliphatic
hydroxyl groups would form a more stable alkoxide bond but would also
hydrolyze in water. Hydrolytically stable bonds could be formed between
lignin and the silicate network through 1,2-benzenediol (13a) and 2-hydroxy-
benzoic acid (13b) groups, which can form the divalent linkages to silicon
shown in equation 13. Since there are 12 benzenediol groups per 100 repeat
units in kraft pine lignin, water-stable bonds can form between kraft lignin
and silicate.
A critical provision needs to be emphasized in this discussion on stability
of ligninsilicate structures, however. The discussion on alkoxidesilicate
bond stability was prefaced by in isolation, meaning a single bond in solu-
tion or a swollen gel. If the alkoxidesilicate bond is formed in a sample which
ultimately loses all water, the sensitivity of the carbonoxygensilicon bond to
hydrolysis, while it still exists, is irrelevant because no water will ever reach
that bond to lyse and separate it. Once a collapsed, hydrophobic aggregate of
silicate and ligninsilicate bonds forms, the capacity of water to penetrate into
the hydrophobic lattice from its surface, break bonds in the lattice, separate
parts of the lattice fromone another, and do these steps on any reasonable time
scale, such as the human lifetime, drops essentially to zero. Unless the
entire aggregate is dissolved in strong base; a dried, aggregate of silicate and
266 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
ligninsilicate bonds will remain stable and combined despite the fact that the
majority of its individual bonds are hydrolyzable.
The products claimed by Blount
[180]
are actually chemisorbed aggre-
gates of lignin in a silicate network with the degree of bonding and dispersion
in the lignin depending on the concentration of lignin in the silicate and the
presence or absence of water in the reaction mixture. Blount formed a lignin
silicate adduct by baking a ground mixture of 50 to 75 weight percent lignin,
25 to 35 weight percent silica gel, and 2.5 to 15 weight percent strong base
together for an hour or by combining these reagents in water and heating to
dryness. The solution reaction will give a more dispersed and uniform prod-
uct, all of the products will have some carbonoxygensilicon bonds, and all
will be increasingly stable with higher uniform dispersion of lignin in the
silicate and more complete dehydration of the formed aggregate. None of
these products have a specic chemical formula or structure, none merit the
nomenclature applied to them by Blount, and none have any commercial use
in 1997.
Phosphorus can be added to lignin by reaction of an alkali lignin with
phosphorus halides, oxyhalides, thiohalides, oxides, suldes, and trivalent
phosphorus esters or amides.
[183]
A representative reaction
[184]
is shown in
equation 14. The action of these strong electrophiles is to halogenate alkyl
groups at the hydroxyl substituent and produce a phosphorus bound to
oxygen, as shown in equation 14a. The reagents can only attack aromatic
hydroxyl groups to form a phosphate ester,
[184]
as shown in equation 14b.
As implied by equation 14b, a polymer with multiple hydroxyl groups
will react with pentavalent or trivalent phosphorous electrophiles to crosslink
the polymer if a molar excess of the phosphorus compound is present.
This was veried by Doughty, who found that phosphorous halide or sul-
de/lignin reactions with a greater than 1/1 mole ratio produced insoluble
products.
[183]
If the number of moles of phosphorous compound are equal or
less than the moles of lignin in the reaction, the soluble reaction product can
MODIFICATION OF LIGNIN 267
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
be further reacted with mono and polyhydroxy compounds to produce ame
retardant resins with lm-forming capacity.
[185]
These types of products, both
soluble and insoluble, can be produced in a number of ways
[186192]
with
degrees of oxygen and alkyl group loss increasing with the severity of the
reaction conditions. However, there is no current use for these compounds
and no production of them. Further, if these materials are to be used as ame
retardants, the halogenation that is coincident with phosphorylation by
strong electrophiles will have to be avoided and other, potentially less cost-
eective, synthetic routes to these materials must be developed.
10. Grafting by Free Radical Polymerization
Once lignin is separated from other plant products, it can be grafted.
Extensive studies on the modication of lignin by graft copolymerization
have been made
[193]
because of the enormous mass of kraft lignin produced
each year by the pulp and paper industry. Graft copolymerization sharply
changes the properties of lignin and allows useful products to be made from
this underutilized portion of biomass.
[194]
Lignin has been grafted with ethenylbenzene (Refs. [195,196]; styrene),
4-methyl-2-oxy-3-oxopent-4-ene (Refs. [197,198]; methylmethacrylate), 2-pro-
penamide (acrylamide), 2-propene nitrile (Ref. [199]; acrylonitrile), cationic
monomers, anionic monomers, and propenoic acid ethoxylates. An index of
compounds listing structure, product name, and trivial name is given in
Table 7. Two types of methods, radiation or chemical, have been used to
attach sidechains of these repeat units to lignin. The radiation methods have
268 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
used both electromagnetic and particle radiation to produce grafting. Low-
energy, electromagnetic irradiation based on visible or ultraviolet light relies
on exciting or decomposing a particular bond either in lignin or in an initiator
present in the reaction mixture. This method, photoinitiation, has not been
used to graft lignin. High energy radiation grafting using either electromag-
netic or particle beams proceeds by ionization and excitation reactions that
produce anionic, cationic, and free-radical sites. Radiation-initiated, ionic
grafting reactions have not been conducted on lignin and therefore, only
free radical polymerization is known to contribute to grafting. The frequency
of active site formation can be expressed as a G value, where G is the
number of molecules reacted or produced by 100 eV of absorbed energy.
Lignin is quite stable to ionizing radiation
[200,201]
having a G
R
value of 0.6
to 0.718. This stability makes lignin a poor candidate for radiation grafting
since in the presence of neat monomer or in solution, initiation will occur far
more readily to form homopolymer than it will to form graft copolymer.
Homopolymer contains only polymerized monomer. Because of these
MODIFICATION OF LIGNIN 269
Table 7. Trivial Names for Compounds and Copolymers
Name Trivial name Structure
poly(1-phenylethylene) polystyrene
ethenylbenzene styrene
poly(1-(1-oxo-2-oxypropyl)-
1-methylethylene)
poly(methylmethacrylate)
2-oxy-3-oxo-4-
methylpent-4-ene
methylmethacrylate
poly(1-amidoethylene) polyacrylamide
2-propenamide acrylamide
poly(1-cyanoethylene) polyacrylonitrile
2-propene nitrile acrylonitrile
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
deciencies, radiation grafting of lignin will not be discussed further. If addi-
tional information on radiation grafting is needed, references are provided
above to radiation grafting studies. Those are the references cited in the list
of monomers grafted to lignin.
The major deciency of radiation grafting is production of homopolymer
instead of graft copolymer. Some reduction in the amount of homopolymer
produced can be achieved by initiating the grafting reaction by chemical
methods. Indeed, Stannett showed that chemical initiation of grafting is
over 8 times more eective in converting monomer to graft copolymer
than is radiation when both are applied to hydrochloric acid lignin. He
notes
[196]
that chemically initiated grafting at 60

C was more eective


than radiation-induced grafting at room temperature. Chemical initiation
can be applied in two ways. First, a reagent which attacks functional groups
on the lignin backbone to produce a grafting site can be used as a grafting
initiator. Alternatively, a reagent which reacts with lignin to form a reactable
functional group is used to form a derivative of lignin. The added groups are
structures such as peroxide or ethene bonds and are then treated or reacted to
initiate grafting. The chemical method can be used to initiate all polymeriza-
tion reactions that do not require a solid supported catalyst
[202]
but only step
and free radical chain reactions have been conducted on lignin.
Unfortunately, for most of the products reported in the literature,
neither of these methods have been used to make the material claimed to
be a graft copolymer. If the polymerization reaction does not start o of the
material to be grafted, then almost all of the monomer is polymerized into
homopolymer with no lignin in the chain. This is a polymerization in the
presence of and it wastes the monomer that was supposed to be used in a
grafting reaction. Thus, a key and often overlooked point in conducting
grafting reactions is to insure that the initiation of the polymerization
occurs on the backbone to be grafted. This requires a special chemistry to
initiate the polymerization. Polymerization methods without such chemistries
merely make homopolymer.
Meister et al. have achieved the grafting of lignin-containing materials
by developing an initiation system that preferentially attacks repeat units in
lignin to create a site for polymer chain growth. The reaction appears general
and works on almost all ethene monomers. By this reaction, the researchers
convert lignin into process polymers for industrial use or thermoplastics for
use in consumer items. Figure 5 is a general diagram showing the materials
reacted with lignin and the applications for the resulting products.
A typical grafting reaction mixes monomer in nitrogen-saturated,
organic or aqueous/organic solvent containing lignin, calcium chloride, and
a hydroperoxide.
[203]
A research group headed by Meister has developed this
reaction chemistry over the past 15 years and have shown that it almost quan-
titatively converts lignin into graft copolymer. The group rst synthesized,
characterized, and tested a spectrum of water soluble, lignin copolymers
270 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
that were nonionic, anionic, or cationic
[204]
and then showed that the water
soluble polymers were eective dispersing, occulating, and surface active
agents. The nonionic polymers and their hydrolysis products are eective
thinners and suspending agents for drilling mud formulations,
[204]
as shown
by the data of Table 8. These test samples compare poly(lignin-g-((1-ami-
doethylene)-r-(1-carboxylatoethylene)), poly((1-amidoethylene)-r-(1-carbox-
ylatoethylene)), and chrome lignosulfonate as aqueous drilling mud
dispersants. The copolymer performs as well as the homopolymer and is
more thermally stable than the lignosulfonate. The anionic polymers products
are thickening agents for uid owcontrol,
[205]
as shownby the data of Table 9.
The high limiting viscosity numbers of these copolymers cause rapid viscosity
increase in water as a function of copolymer concentration. The cationic poly-
mers products are dewatering aids for sewage treatment,
[206]
as shown by the
data of Table 10.
The ethoxylate esters of propenoic acid are useful as prepolymers for
urethane formation but these materials are not as eective as those prepared
by Wolfgang Glassers research group. Therefore, only Dr. Glassers prod-
ucts will be discussed.
In reactions with ethenylbenzene, lignin was used to make thermoplastic
materials. Data for a spectrum of reactions run to optimize yield and create
samples of dierent molecular weight and composition
[207]
are given in
Table 11. Samples PE 1 to PE 10 were made with ethenylbenzene as monomer.
These products have been shown to be poly(lignin-g-(1-phenylethylene))-
containing materials by a series of solubility and extraction tests and are
MODIFICATION OF LIGNIN 271
Figure 5. Monomers that may be grafted onto lignin and uses of the resulting products.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
272 MEISTER
Table 8. Properties of Test Muds Before and After Hot Rolling
a
Property
Base Mud
Graft
Copolymer
Fraction
Homopolymer
Fraction
Chrome
Lignosulfonate
Before After Before After Before After Before After
Viscosity in centipoise at a shear rate of:
1020 s
1
52 69 74 42 74 42 29 50
510 s
1
36 47 49 24 49 24 16 33
340 s
1
30 29 40 17 40 17 12 26
170 s
1
21 28 27 10 27 10 7 18
Gel strength in lb/100 ft
2
mud has set for:
10 sec 6 12 5 3 5 2 2 11
10 min 25 35 20 3 20 3 9 24
Apparent Viscosity (cp) 26 33 37 21 37 22 15 25
Plastic Viscosity (cp) 16 22 25 18 25 19 13 17
Yield Point (lb/100 ft
2
) 20 25 24 6 24 5 3 16
API Filtrate Volume (mL) 12.0 13.8 7.8 8.0 7.8 8.4 11.6 14.0
pH 9.1 8.6 9.0 8.0 9.0 8.0 9.5 8.2
High Pressure,
High Temperature
Filtrate (mL)
62.8 52.4 46.0 64.0
a
Hot rolling is agitating the mud at 121

C for 16 hours.
Table 9. Synthesis Data and Physical Characteristics of Graft Terpolymer
a
Sample
Number
Reactants
2-Propen
amide (g) A (g)
Dimethyl
sulfoxide
(mL)
CaCl
2
(g) B (g)
Yield
(wt%)
[]
(dL/g)
1 1.60 4.66 20 0.50 0.15 70.12 10.52
2 1.60 4.66 50 0.50 0.15 86.98 11.40
3 1.60 4.66 50 0.50 0.25 78.40 7.40
4 1.60 4.66 40 0.50 0.40 69.82 9.30
5 1.60 5.16 30 0.50 0.15 78.79 12.59
6 1.60 5.16 30 0.50 0.15 77.27 6.81
7 1.60 4.66 30 0.50 0.15 87.28 10.46
8 2.58 1.87 30 0.50 0.39 67.89 0.953
9 2.56 1.86 30 0.53 0.39 79.49 2.46
10 21.98 15.99 219 4.35 3.35 91.02 1.97
a
All reactions, save #10, contained 0.50 g of lignin and 0.15 mL of cerium (4)
ion. Reaction #10 contained 4.39 g of lignin and 1.28 mL of cerium ion.
(A) 2,2-Dimethyl-3-imino-4-oxohex-5-ene-1-sulfonic acid.
(B) Hydroperoxide. Samples 1 to 7: Values are weight of 1,4-dioxa-2-hydroper-
oxycyclohexane in g. Samples 8 to 22: Values are amount of aqueous solution of
1,2-dioxy-3,3-dimethylbutane in mL. Equivalents/mL7.23 10
3
.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
formed with 90% or more grafting eciency for lignin. These materials have
been shown to be thermoplastics,
[207]
coupling agents for wood and plastic,
[208]
and biodegradable plastics.
[209]
The thermoplasticity of the graft copolymers can be veried by measure-
ments of the glass transition temperature of the newsolids. The glass transition
temperature is the temperature at which an amorphous solid becomes ductile
and is a characteristic of thermoplastic materials. Samples of 5 to 10 mg of
reaction product were heated at 10

C per minute in a dierential scanning


calorimeter to monitor heat capacity as a function of temperature. The tem-
perature of each transition produced by each copolymer is shown in Table 12.
These products also occupy surfaces on wood and act to alter the wetting
properties of the plant material. The lignin and copolymer solutions gave
smooth, adherent surface coatings on the wood with contact angles against
water of 90 to 110

while plastics and plasticlignin mixtures did not give


adherent coatings. These data show that copolymers of lignin are surface
active, preferentially orienting the lignin portion of the product towards
wood while the plastic sidechain is oriented outward to create a new surface
with dierent wetting properties. Thus, these copolymers are surface-active,
MODIFICATION OF LIGNIN 273
Table 10. Synthesis and Application Data for Catinoic Graft Copolymer
A: Synthesis Data
Reactant
Sample
Number
Lignin
(g)
CaCl
2
(g)
A
(g)
C
(g)
DMSO
(g)
E
(mL)
Yield
%
1 0.50 0.93 1.92 5.15 30.72 0.50 80.18
2 0.50 0.99 1.29 8.36 29.28 0.50 70.44
3 0.50 1.02 0.64 10.21 31.40 0.50 91.46
4 0.50 1.07 12.77 30.71 0.50 70.91
B: Dewatering Data of Cationic Lignin Graft Copolymer with Methylsulfate Anion
Sample
Number
Concentration
in Sludge (ppm)
Filtrate Volume After Being on the Filter
for the Given Number of Minutes
1 2 3 4 5 10
BLANK 00.0 <10 10
1* 300 35 45 55 60 70
1* 450 30 45 55 65 70
3 150 28 46
4 150 52 74
CommPol. 150 54 72
a
a
Performance data for the commercial polymer in use at the Detroit Water and Sewerage
Works in 1991. (A) 2-Propenamide. (C) Cationic monomer. 2-Methyl-N7,N7-trimethyl-7
ammonium-3-oxo-4-oxyoct-1-ene chloride or 2-methyl-N7,N7-trimethyl-7-ammonium-3-
oxo-4-oxyoct-1-ene methylsulfate. (E) 30% Hydrogen peroxide (equivalent weight: 8.383
meq/mL). *Cloudy ltrate.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
coupling agents which can bind wood to hydrophobic phases such as plastic.
This coupling process works best when the wetting agent has been synthesized
so that the sidechain attached to the lignin during the preparation of
the macromolecular, surface active agent is chemically identical to the plastic
hydrophobic phase that is to be bound or connected to the wood. Thus, to bind
poly(1-phenylethylene) [Trivial name polystyrene] to wood, coat the wood
with poly(lignin-g-(1-phenylethylene)) and to bind poly(1-cyanoethylene)
274 MEISTER
Table 11. Copolymerization Reactions of Lignin and Ethene Monomers
Reactants (g)
Sample
Number Lignin Monomer CaCl
2
H
2
O
2
(mL) Solvent
Yield
(g/wt.%)
PE 1 2.00 18.76 2.02 1.0 20.04 17.80/85.7
PE 2 2.01 18.78 2.02 5.0 20.02 18.53/89.1
PE 3 3.03 18.78 2.00 2.0 20.00 19.14/87.8
PE 4 2.00 18.76 1.01 2.0 20.10 18.84/90.8
PE 5 2.01 4.69 2.04 2.0 20.01 5.68/84.8
PE 6 2.01 9.39 2.02 2.0 20.00 10.42/91.4
PE 7 2.01 14.07 2.03 2.0 20.10 14.95/92.8
PE 8 2.02 23.45 2.04 2.0 20.07 23.76/93.3
PE 9 8.00 28.15 8.00 8.0 40.02 33.16/91.7
PE 10 8.04 18.76 8.00 8.0 40.03 24.14/90.1
CY-AM 1 4.10 9.35/0.0* 3.05 4.00 20.04 13.86/103.0
CY-AM 2 0.50 0.35/4.17 0.5 0.53 28.91 4.82/96.4
CY-AM 3 0.50 0.78/3.74 0.5 0.53 28.91 4.91/97.8
CY-AM 4 0.50 1.19/3.33 0.5 0.53 28.91 3.50/69.7
CY-AM 5 0.50 0.79/3.75 0.5 0.53 28.91 3.28/65.1
CY 1 4.10 9.35/0.0 3.05 4.0 20.04 13.86/103.0
CY 2 3.95 6.08/0.0 3.07 4.0 20.07 9.83/98.0
CY 3 4.05 3.13/0.0 3.02 4.0 20.57 6.34/88.3
CY 4 4.02 6.15/0.0 3.05 4.0 25.02 10.14/99.7
CY 5 4.00 6.16/0.0 2.51 5.0 20.01 9.29/94.3
MBuD 1 4.11 9.32 3.02 4.0 20.29 4.21/31.3
MBuD 1 4.01 3.14 3.05 4.0 20.32 4.39/61.4
MBuD 1 3.98 6.27 3.07 4.0 21.36 3.71/36.2
MPrPe 1 4.00 17.17 3.03 4.48 20.04 19.54/92.30
MPrPe 2 4.03 11.57 3.09 4.48 20.11 15.29/98.01
MPrPe 3 3.97 6.37 3.06 4.48 20.27 9.69/93.71
MPrPe 4 4.00 11.78 2.98 4.48 20.06 16.2/102.67
MPrPe 5 4.05 11.53 2.57 5.60 20.35 15.69/100.71
MPrPe 6 8.00 6.1 3.03 4.70 30.42 13.53/95.96
MPrPe 7 8.1 2.41 3.08 4.48 30.21 9.86/93.82
MPrPe 8 7.99 17.52 3.11 8.96 30.23 24.22/94.94
MPrPe 10 2.00 18.07 2.06 4.48 20.11 20.63/102.79
*First number weight of 2-propene nitrile added and second number weight of
2-propenamide added.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
[Trivial name polyacrylonitrile or orlon] to wood, coat the wood with
poly(lignin-g-(1-cyanoethylene)).
The coupling of wood by a grafted plant part to a plastic, that has the
same or similar composition as the sidechain on the plant part, increases
the binding strength of the plastic
[210]
to the wood. This was proven by
performing lap shear tensile strength tests on birch strips onto which were
injection molded blocks of plastic. The samples with a grafted product
coating with the same repeat units in the sidechain as in the plastic gave
20 to 50 percent higher tensile strength. The coupling experiments were
performed as follows. Birch tongue depressors, a medical product of
1.75 mm thickness were cut into suitable sizes to match an injection
mold. Kraft pine lignin was reacted into a graft copolymer as previously
described. The homopoly(1-phenylethylene) used in coupling tests is a
recovered fraction of the reaction product of mechanical pulp and ethenyl-
benzene. Coatings of copolymer or any comparison mixtures or control
blanks were prepared as a 10 weight percent solution in dimethylformamide
and spread on the wood. The plastic phase was Amoco RIPO, from Amoco
Chemical Company. Injection molding was done on a Milberry, Model 50
Mini-Jector. Experimental conditions were: cylinder temperature, 288

C;
nozzle temperature, 172176

C; pressure, 3450 kPa; pressure holding time,


12 seconds; and chilling time, 12 minute.
Lap shear strength of the pieces of wood with plastic injection molded
to them
[211]
was tested on a Instron, Model 4200, Universal testing instru-
ment USDA. Experimental conditions were: room temperature, 23

C; room
relative humidity: 50 percent; crosshead speed, 2.54 mm/min; with the sample
MODIFICATION OF LIGNIN 275
Table 12. Dierential Scanning Calorimetry Data for Lignin, Poly(1-phenylethylene),
and Graft Copolymer
Sample
Number
Lignin
Percent in
Reaction
Lignin
Percent in
A Peak(s,

C)
Ramp
(

C/min)
Amoco RIPO#
(pure poly(1-phenylethylene))
0 0 102.6 10
Kraft Pine Lignin
(pure lignin)
100.0 100.0 116.17130.09 10
Copolymer 1 9.6 10.3 94.82 (114.62)* 10
Copolymer 2 22.0 27.3 98.43 (133.97)* 10
Copolymer 3 30.0 32.2 98.23 (124.10)** 10
Copolymer 4 30.0 34.5 102.35 (144.48)** 20
Copolymer 5 30.0 32.3 95.73 133.25 10
Copolymer 6 46.0 50.5 94.11 125.12 10
Copolymer 7 46.0 51.8 101.63 143.27 20
(A) Copolymerization reaction product. *Very small peak. **Small peak. # A commer-
cial poly(1-phenylethylene) from Amoco Chemical Company.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
in hand-fastened grips and an aluminum specimen holder. The lap shear
strengths of the woodplastic samples are summarized in Table 13. The
copolymer samples were fractionated by benzene extraction. The reaction
product was labeled Product A. The benzene-soluble extract of the product
became Ben. Ex. while the benzene insoluble portion of the product was
labeled Product B.
In almost all cases, coating the wood with any of the three fractions of
the graft copolymer of lignin and ethenylbenzene (Product A, Product B, and
Product Ben. Ex.) provides stronger adhesion between wood and commercial
poly(1-phenylethylene) than coating the wood with mechanical mixtures,
pure poly(1-phenylethylene), pure lignin, or nothing (blank). The data
show the copolymers to be eective coupling agents.
White rot Basidiomycetes were able to biodegrade graft copolymers of
lignin and ethenylbenzene containing dierent proportions of lignin and
276 MEISTER
Table 13. Summarized Adhesion Strength Results
Coating Material Adhesion Strength (kPa)
30-151-2A (10.45% lignin) 2422.1 219.3 (3)*
30-151-2B 2313.9 488.2 (3)
30-151-2Ben.Ex. 2126.4 44.1 (3)
10% Lig. 90% PS 2209.1 251.0 (5)
35-120-1A (24.23% lignin) 2313.9 81.4 (3)
35-120-1B 2278.7 294.4 (2)
35-120-1Ben.Ex. 2094.0 213.7 (3)
24% Lig. 76% PS 2027.1 185.5 (4)
35-110-3A (32.17% lignin) 1930.5 304.1 (3)
35-110-3B 1911.2 184.8 (3)
35-110-3Ben.Ex. 2670.4 207.5 (3)
32% Lig. 68% PS 1949.2 265.4 (5)
35-115-3A (51.70% lignin) 2838.6 60.0 (3)
35-115-3B 2723.4 328.9 (3)
35-115-3Ben.Ex. 2707.6 70.3 (3)
50% Lig. 50% PS 1843.0 91.0 (5)
Poly(1-phenylethylene) (PS) 2040.9 206.8 (5)
10% Lig. 90% PS 2209.1 251.0 (5)
24% Lig. 76% PS 2027.1 185.5 (4)
32% Lig. 68% PS 1949.2 265.4 (5)
50% Lig. 50% PS 1843.0 91.0 (5)
Lignin 2123.6 164.1 (5)
Blank (treated with DMF) 2022.2 120.7 (2)
Blank (treated with nothing) 1825.7 165.5 (4)
*Number of valid repetitions of the tensile strength test in parentheses. Lig. Lignin.
PSpolystyrene. DMFdimethylformamide.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
poly(1-phenylethylene).
[212]
The biodegradation tests were run on lignin/ethe-
nylbenzene copolymerization products which contained 10.3, 32.2, and 50.4
weight percent lignin. The polymer samples were incubated with white rot,
lignin-degrading organisms Pleurotus ostreatus, Phanerochaete chrysosporium,
Trametes versicolor, and brown rot, cellulose-degrading organism
Gleophyllum trabeum. Over a 68 day period, white rot fungi degraded the
plastic samples at a rate which increased with increasing lignin content in
the copolymer sample. Both poly(1-phenylethylene) and lignin components
of the copolymer were readily degraded. Pure poly(1-phenylethylene) pellets
were not degradable in these tests. Observation by scanning electron micros-
copy of incubated copolymers showed a deterioration of the plastic surface.
Brown rot fungus did not aect any of these plastics. White rot fungi produced
and secreted oxidative enzymes associated with lignin degradation in liquid
media during incubation with lignin-poly(1-phenylethylene) copolymer. All of
these applications represent signicant markets for modied lignin.
In examples Cy-Am 1 to 5 of Table 11 the monomers used were
2-propene nitrile [107-13-1] and 2-propenamide [79-06-1]. In examples Cy
1 to 5 of Table 11, the monomer used was 2-propene nitrile [107-13-1]. The
compounds in the rst group are water absorbing agents while those in the
second group are thermoplastics and biodegradable plastics.
[213]
In examples
MBuD 1 to 3 of Table 11, the monomer used was 2-methyl-1,3-butadiene
[78-79-5]. These materials are uncrosslinked elastomers and potential rubber
additives. In examples MPrPe 1 to 9 of Table 11, the monomer used was
2-methyl-2-oxy-3-oxopent-4-ene [80-62-6]. These materials are thermoplas-
tics and biodegradable plastics.
[214]
11. Grafting by Anionic Chain Polymerization
Lignin can be grafted with alkane epoxides to form polyether adducts
and this reaction has been extensively studied by the research group of
Dr. Wolfgang Glasser at Virginia State University and Polytechnic
Institute. The reaction is initiated by base attack on the phenolic hydroxyl
groups to form phenoxide groups. The phenoxide groups then attack the
polar epoxide ring. A signicant side reaction in this ring opening polymer-
ization is a nucleophilic attack on the ring by hydroxyl anion. These reactions
are shown in Figure 6.
Alkoxylation is critically important as a precursor reaction which
changes both the physical properties of the lignin and its chemistry. The
most important physical property changed is the glass transition temperature
of the lignin. As the weight fraction of ethylene or propylene oxide in the
product increases, the glass transition temperature of the product falls. Since
a reduction in T
g
is synonymous with a lowering of viscosity at any tempera-
ture and a lowering of the temperature at which ow starts, this change
MODIFICATION OF LIGNIN 277
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
makes lignin a owable liquid at temperatures 100 to 200

C below the usual


temperatures for viscous lignin ow.
Simultaneously, Figure 6 shows that alkoxylation leaves lignin
capped with chains ending in primary hydroxyl groups. These alcohols
are more reactive in some chemistries than the phenolic hydroxyl groups
that they have capped. A particularly important example of one such
chemistry is the reaction of the alkoxylated lignin with an organic isocyanate
or diisocyanate. The reaction of the alkoxylated lignin with an isocyanate-
terminated polymer allows Glasser et al. to add polymeric sidechains of
cellulose
[215]
or polycaprolactam
[216]
to the lignin graft copolymer.
Reactions of alkoxylated lignin with diisocyanates produce thermoset
materials because the lignin polyol is always polyfunctional with a function-
ality greater than 2. The isocyanate-alcohol reaction produces a urethane
linkage that, when repeated, creates a crosslinked, non-reformable polyur-
ethane. This is shown in Figure 7. Abroad spectrumof lignin-based urethanes
have been made and tested. The data show that these materials match if not
exceed the properties of synthetic polyurethanes made without lignin.
[217]
The alkoxylated lignin requires an isocyanate to hydroxyl group ratio of
1.5 to form eective networks. This is a signicantly higher ratio of expensive
isocyanate to hydroxyl units than is used in commercial practice. The work of
Glasser et al. shows that the glass transition temperature, T
g
, of a alkoxylated
lignindiisocyanate network depends on the weight percent lignin in the net-
work, the chemical structure of the alkoxide, and the chemical structure of
the diisocyanate.
[217]
As lignin content increases, T
g
of the network increases.
Networks made with oxyprop-1,2-ylene repeat units, OCH
2
CH(CH
3
), have
278 MEISTER
Figure 6. Alkoxylation of a lignin repeat unit.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
lower T
g
s than networks made with oxyeth-1,2-ylene repeat units when the
diisocyanate used in the network is 1,6-diisocyanaohexane, an aliphatic
diisocyanate. When the diisocyanate is aromatic, however, there is no dier-
ence between the T
g
s of the materials made with oxypropyl or oxyethyl
repeat units. Networks made with aliphatic diisocyanate generally have
lower T
g
s than networks made with aromatic diisocyanate.
The mechanical properties of these networks, modulus of elasticity and
ultimate strength, are dependent on the method of preparing the network.
[217]
While the chemical composition and repeat units of the network are impor-
tant factors that inuence the properties of the crosslinked product, the most
important inuence on physical properties of the network is the way the
polymeric, alkoxide chain is introduced into the network. Products made
by adding a pure polyether that has not been grafted onto lignin have
higher modulus of elasticity and ultimate strength than products made
with alkoxylated lignin copolymer. This dierence in physical properties
probably reects the steric hindrance and reduced diusion produced by
grafting a polyether onto lignin. The copolymer produces a less uniform
distribution of crosslinks and a greater segregation of network parts than
does the blend of pure polyether with the other reagents.
A key feature in forming this material that is not widely recognized out-
side of the polyurethane industry is that the reactants must be uids at 60

C,
the common temperature for conducting this reaction. This is why the reduc-
tion in glass transition temperature with increasing degree of alkoxylation in
lignin is so important. By decreasing the temperature at which modied lignin
will ow, the alkoxylation reaction allows lignin to not only be a uniformly
reactive, poly primary alcohol, but also a uid polyol that can be processed by
MODIFICATION OF LIGNIN 279
Figure 7. Reaction of ethoxylated lignin with a diisocyanate.
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
procedures such as reaction injection molding into thermoset solids. The pros-
pects for application of this chemistry to foams, networks, and consumer
goods is great since the materials match the properties of existing products,
the area of application is high-prot-margin, specialty chemicals, and lignin
sharply reduces cost by replacing more expensive polyol.
III. CONCLUSIONS
The utilization of lignin will increase in the next quarter century as
demand for aromatic carbon exceeds the supply available from a decreasing
inventory of oil. Lignin is a more homogeneous material than petroleum that
can provide a relatively uniform supply of alkyloxyaromatics for decomposi-
tion, utilization as extracted from the plant, and use as a backbone to create
larger, altered polymers. The market that not only has the highest rate of
growth in 1998 but also promises the largest increases in utilization of lignin
is the use of lignin as an adhesive in wood composites. Lignin constitutes 17
weight percent of the solids in most exterior-grade plywood and will become
a progressively larger fraction of the binder in laminates and ber, strand, or
wafer board. Lignin has a potential to become a functional photostabilizer
and free radical trap because of its high molar absorptivity at ultraviolet
wavelengths below 300 nm and its ability to trap and maintain free radicals.
However, these applications will require careful formulation of a product
that will include a chemically altered lignin in place of the product that
can be extracted from the plant. A possible use for lignin in the future is
the addition of alkali lignins to pet and human food as roughage, a ber
source, or a cancer protection agent.
Thermal or chemical decomposition of lignin to produce oxyaromatics
will grow when the depletion of low cost petroleum becomes pronounced
between 2025 and 2035 A.D. Examples of this technology would be nonionic
surfactants containing retorted, ethoxylated lignin or cresylic acid produced
by pyrolysis. While this is a market for a small mass of lignin, it is a high
prot margin market with specialty chemical applications stretching from
adhesives to xylene manufacture.
The commonmodicationreactions of ligninare alkylation, dealkylation,
sulfomethylation, methylolation, sulfonation, amination, nitroxide formation,
carboxylation, acylation, silylation, halogenation, nitration, phosphorylation,
hydrogenolysis, grafting, and oxyalkylation. Of these reactions, nitroxide
formation, carboxylation, acylation, silylation, halogenation, nitration, and
phosphorylation are laboratory processes that have no commercial applica-
tion in 1998. Alkylation and dealkylation are common side reactions during
lignin extraction and recovery. Alkylation forms a lignin thermoplastic but this
brittle material has no current utility. Sulfomethylation and sulfonation form
lignosulfonates which have a broad market as tanning agents, electrode
280 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
stabilizers, suspending agents, and dispersing agents. They are the most widely
used, modied lignin. Sulfomethylation is only used when an increased content
of sulfonate groups is needed in the lignin product. It is applied to kraft lignins
by the Westvaco Corporation to formdye dispersants that are marketed under
the Reax trade name. The current markets for lignosulfonates, a moisture
retention agent in cement grouts, a dust suppressor for road treatments, an
additive for battery electrodes, or a thickening agent in inks, are small but
stable markets. Methylolation is a pre-reaction to prepare lignins for use as a
phenol extender in phenolmethanal resins. It is used commercially in the
forest products industry. Amination is used to form lignin with beta keto
amine groups in it. This amine lignin is used to formthermally stable, aqueous
emulsions of asphalt for use in road repair.
Graft copolymerized lignin is a research material being tested for indus-
trial and consumer applications. Copolymerization of lignin with polar
monomers to create process polymers will permit the use of lignin in water
treatment, sewage dewatering, thickening, and dispersion. These nonionic,
anionic, and cationic, water soluble polymers will be industrial process poly-
mers which allow the production of consumer and manufactured products.
Copolymerization of lignin with nonpolar monomers to create thermoplas-
tics will permit the use of lignin in the commodity plastics market. These
materials are biodegradable thermoplastics, thermoplastic composites, and
coupling agents to incorporate wood and plastic into a single phase.
The alkoxylation of lignin will permit its use in processes requiring a
uid reagent for further modication. The largest market for this new mate-
rial will be the production of urethane solids and foams from the reaction of
the alkoxylignin with a diisocyanate. This high value material will have
applications as an engineering plastic or as insulation. Alkoxylated lignin is
also useful as an intermediate to be grafted into new materials by reaction
with step-reaction polymers terminated with a group that creates covalent
bonds with hydroxyl groups. While carboxylic acid-terminated, step poly-
mers can be reacted with alkoxylignin to form alkoxylignin esters, thermo-
dynamic and equilibrium forces dictate that the best products from this
chemistry will come from step-reaction polymers capped with isocyanate
groups. New products of alkoxylated lignin covalently bonded to cellulose
acetate or caprolactam are known and are available for utilization.
QUESTIONS
1. What is the source of lignin?
2. Why is lignin made and to what use is it put when it is rst manufactured?
3. What are the three monomers that are used to make lignin?
4. Name three methods of recovering lignin once it has been synthesized.
5. Name three products from the pyrolysis of lignin.
MODIFICATION OF LIGNIN 281
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
6. Describe the use of lignin in the production of Bakelite polymers,
phenolmethanal, network polymers.
7. What are the potential uses for lignin as a food additive?
8. Name two of the three common methods for adding alkyl groups to
lignin.
9. A common chemical reaction based on acid catalyzed addition of
methanal and a primary or secondary amine is conducted on alkali
lignin in aqueous suspension to convert it into a commercial product.
What is the name of this reaction, what product is formed by it, and
what is the use of that product?
10. The study of lignin sulfonation dates to 1866. Why does this reaction
have suchalonghistoryof extendedinvestigations andfor what is it used?
11. What is a graft copolymer of lignin and how is it made?
REFERENCES
1. Henry I. Bolker. Natural and Synthetic Polymers. An Introduction; ISBN
0-8247-1060-6; Marcel Dekker: New York, 1974; p. 580.
2. Eero Sjostrom. Wood Chemistry, Fundamentals and Applications; ISBN
0-12-647480-X; Academic Press, 1981; p. 69.
3. Falkehag, S.I. Appl. Polymer. Symp. 1975, 28, 247257.
4. Kent Kirk, T.; Higuchi, T.; Chang, H. Lignin Biodegradation: Microbiology,
Chemistry, and Potential Applications; ISBN 0-8493-5459-5; CRC Press, 1980;
Vol. 1, p. 5.
5. Fengel, D.; Wegener, G. Wood, Chemistry, Ultrastructure, Reactions; Walter
de Gruyter & Co.: Berlin, 1989; pp. 133139.
6. March, J. Advanced Organic Chemistry, 4th Ed.; J. Wiley & Sons: New York,
1992; pp. 3038.
7. Meshitsuka, G.; Isogai, A. In Chemical Modication of Lignocellulosic Materi-
als; David N.-S. Hon, Ed.; Marcel Dekker Inc.: New York, 1996; pp. 1134.
8. Wolfgang G. Glasser.; Heidi R. Glasser. The Evaluation of Lignins Chemical
Structure by Experimental and Computer Simulation Techniques. Paperi ja
Puu 1981, (#2), 7183.
9. Wolfgang G. Glasser.; Heidi R. Glasser; Morohoshi, N. Simulation of
Reactions with Lignin by Computer (SIMREL). 6. Interpretation of
Primary Experimental Analysis Data. Macromolecules 1981, 14, 253262.
10. Erickson, M.; Larsson, L.; Miksche, G.E. Acta Chem. Scand. 1973, 27,
16731678.
11. Nimz, H. Angew. Chem. Int. Ed. Engl. 1974, 13, 313321.
12. Oxacyclopentene units are contained in 2,3-dihydrobenzofuran structures.
13. Bjorkman, A.; Person, B. Svensk Papperstindn. 1957, 60, 285.
282 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
14. Sarkanen, K.V.; Ludwig, C.H. Lignins, Occurrence, Formation, Structure, and
Reactions; John Wiley & Sons: New York, 1971; pp. 2223.
15. Freudenberg, K.; Chen, C.-L.; Harkin, J.M.; Nimz, H.; Renner, H. Chem.
Commun. 1965, 1 (#1), 224225.
16. Gellerstedt, G.; Kringstad, K.; Lindfors, E.L. p. 302 of Singlet Oxygen.
Reactions with Organic Compounds and Polymers, Banby, B., Rabek, J.F.,
Eds.; Wiley Interscience: Chichester, 1978.
17. Atalla, R.H.; Agarwal, U.P. Raman Microprobe Optimization and
Sampling Technique for Studies of Plant Cell Walls. In Microbeam Analysis,
Romig, A.D. Jr., Goldstein, J.I., Eds.; San Francisco Press, CA, 1984; pp.
125126.
18. Atalla, R.H.; Agarwal, U.P. Raman Microprobe Evidence for Lignin
Orientation in the Cell Walls of Native Woody Tissue. Science 8 February,
1985, 227, 636638.
19. Wardrop, A.B. In Cellular Ultrastructure of Woody Plants, Cote, W.A., Ed.;
Syracuse University Press, 1965.
20. Pickett-Heaps, J.D. Protoplasma 1968, 65, 181.
21. Wardrop, A.B. Unpublished Results.
22. Rydholm, S.A. Pulping Processes; Interscience Publishers: New York, 1965.
23. Sarkanen, K.V.; Ludwig, C.H. Lignins, Occurrence, Formation, Structure, and
Reactions; John Wiley & Sons, New York, 1971; pp. 639694.
24. Sarkanen, K.V.; Ludwig, C.H. Lignins, Occurrence, Formation, Structure, and
Reactions; John Wiley & Sons: New York, 1971; pp. 597637.
25. Pye, E.K.; Lora, J.H. The Alcell ProcessA Proven Alternative to Kraft
Pulping. Tappi 1991, 74 (#3), 113118.
26. Lora, J.H.; Pye, E.K.; Winner, S.R. Industrial Scale Alcohol Pulping, Forest
Products Symposium, 1990, 3539.
27. Lora, J.H.; Creamer, A.W.; Wu, L.C.F.; Goyal, G.C. Chemicals Generated
During Alcohol Pulping: Characteristics and Application. Sixth International
Symposium on Wood and Pulping Chemistry, 1991, 431438.
28. Bjorkman, A. Svensk Papperstidn. 1956, 59, 477.
29. Pew, J.C. Tappi 1957, 40, 553.
30. Nord, F.F.; Schubert, W.J. Holz Forschung 1951, 5, 1.
31. Nord, F.F.; Schubert, W.J. Tappi 1957, 40, 285.
32. de Stevens, G.; Nord, F.F. Fortschr. Chem. Forsch. 1954, 3, 70.
33. de Stevens, G.; Nord, F.F. J. Am. Chem. Soc. 1951, 73, 4622.
34. Kudzin, S.F.; Nord, F.F. J. Am. Chem. Soc. 1951, 73, 690, 4619.
35. Nord, F.F.; de Stevens, G. Naturwissenschaften 1952, 39, 479.
36. Pew, J.C. J. Am. Chem. Soc. 1952, 74, 2850.
37. Hagglund, E. Cellulosechemie 1923, 4, 84.
38. Sakakibara, A.; Nakayama, N. J. Japan Wood Res. Soc. 1962, 8, 153.
39. Heuser, E.; Skioldebraud, C. Destructive Distillation of Lignin (Trans.), Z.
Angew. Chem. 1919, 321, 41.
40. Fisher, F.; Schrader, H. Ges. Abh. Kennt. Kohle, 1920, 5, 106 (in German).
41. Tropsch, H. (Trans.), Ges. Abh. Kennt. Kohle 1921, 6, 293.
42. Fletcher, T.L.; Harris, E.F. J. Am. Chem. Soc. 1947, 69, 3144.
43. Goheen, D.W.; Henderson, J.T. Cell. Chem. Technol. 1978, 12, 363.
MODIFICATION OF LIGNIN 283
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
44. Nimz, H. In Fourth International Symposium on Wood and Pulping Chemistry;
Paris, 1987; pp. IIIAIIIK.
45. Whittington, L.E.; Naae, D.G.; Davis, C.A.; Templeton, J.E. Presented at the
Symposium on Emerging Materials and Chemicals from Biomass and
Wastes. American Chemical Society National Meeting, Washington, DC,
August, 1990.
46. Naae, D.G.; Whittington, L.E.; Davis, C.A. Presented at the Symposium on
Alteration and Utilization of Lignin. American Chemical Society National
Meeting, San Francisco, CA, April, 1992.
47. Naae, D.G.; Whittington, L.E. Presented at the Kyosti Sarkanen Memorial
Symposium. American Chemical Society National Meeting, Denver, CO,
April, 1993.
48. Maloney, G.T. Chemicals from Pulp and Wood Waste, Noyes Data Corp.:
Park Ridge, NJ, 1978, p. 158.
49. Gendler, J.G.; Huibers, D.T.A.; Parkhurst, H.J. Jr. Presented at the American
Chemical Society Meeting, Kansas City, KS, 9/82.
50. Freudenberg, K.; Adam, K. Ber. 1941, 74, 387.
51. Lindors, T.; Enkvist, T. Finska Kemist. Medd. 1965, 74, 29.
52. Brewer, C.P.; Cooke, L.M.; Hibbert, H. J. Am. Chem. Soc. 1948, 70, 57.
53. Swiss Patent 305,712, 1955.
54. Giesen, J. Process for Cleavage of Lignin to Produce Phenols. U.S. Patent
2,991,314, 1961.
55. Oshima, M.; Maeda, Y.; Kashima, K. Process for Liquefaction of Lignin.
Can. Patents, 700,209 and 700,210, 1964.
56. Goheen, D.W. Hydrogenation of Lignin by the Noguchi Process, p. 205. In
Lignin, Structure and Reactions, Marton, J., Ed.; American Chemical Society,
Washington, DC, 1966.
57. Goheen, D.W. Chemicals from Lignin, Paper Presented to the 8th World
Forestry Congress, Jakarta, Indonesia, 1978.
58. Fisher, J.H.; Marshall, H.B. U.S. Patents 2,576,7523, Nov. 27, 1951.
59. Fisher, J.H.; Sankey, C.A. U.S. Patent 2,576,754, Nov. 27, 1951.
60. Craig, D.C.; Logan, C.D. Can. Patent 615,552, Feb. 28, 1961.
61. Pauly, H.; Feuerstein, K. Ger. Patent 552,887, Sept. 28, 1928.
62. Gohen, D.W. U.S. Patent 2,840,614, 1958.
63. Sarkanen, K.V.; Ludwig, C.H. Lignins; Occurrence, Formation, Structure,
and Reactions, ISBN 0-471-75422-6; J. Wiley, 1971; p. 1.
64. Nims, H.H. Chapter 5: Lignin-Based Wood Adhesives, Wood Adhesives, Pizzi,
A., Ed.; Marcel Dekker Inc.: New York, 1983.
65. Marton, J. Tappi 1964, 47, 713.
66. Gardner, D.J.; McGinnis, G.D. J. Wood Chem. Tech. 1988, 8 (#2), 261288.
67. Muller, P.C.; Glasser, W.G. J. Adhesion 1984, 17, 157.
68. Adams, J.W.; Schoenherr, M.W. U.S. Patent 4,306,999, 1981.
69. Hollis, J.W.; Schoenherr, M.W. U.S. Patent 4,303,562, 1981.
70. Gardner, D.J. Selected Investigations into the Chemistry and Utilization of
Biomass Lignins, Ph.D. Thesis, Forest Products Laboratory, Mississippi
State University, Columbus, MS, DA8609771.
71. Sellers, T. Jr. Panel World 1991, 31 (5), 2629 44.
284 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
72. Lin, S.Y. Dyestus and Dyeing Method Using Lignin Adduct Dispersant.
U.S. Patent 4,492,586, Jan. 8, 1985.
73. DePaoli, M.A.; Furlan, L.T. Polym. Degrad. Stab. 1985, 11, 327337.
74. DePaoli, M.A.; Schultz, G.W.; Furlan, L.T. J. Appl. Polym. Sci. 1984, 29,
2493.
75. Kularatne, K.W.S.; Scott, G. Eur. Polym. J. 1978, 14, 835.
76. Keilen, J.J.; Dougherty, W.K.; Cook, W.R. Ind. Engng. Chem. 1952, 44, 163.
77. Lin, S.Y. Progress in Biomass Conversion, Tillman, D.A., Jahn, E.C., Eds.;
Academic Press, 1983; Vol. 4, 00.3178.
78. Standen, A., Ed. The Kirk-Othmer Encyclopedia of Chemical Technology,
1964; Vol. 3, p. 258.
79. Stewart, A.; Pitrot, A.; Willihnganz, E. U.S. Patent 2,479,983, 1949.
80. Ritchie, E.J. J. Electrochem. Soc. 1953, 100, 5359.
81. Roy, A.K. Indian Pulp Paper (India) 1983, 28 (#2), 1013.
82. Shalizadeh, F., Sarkanen, K.V., Tillman, D.A., Eds., Thermal Uses and
Properties of Carbohydrate and Lignins; Academic Press, 1976; p. 217.
83. Slavin, J.L. Dietary Fiber: Classication, Chemical Analysis, and Food
Sources. J. Am. Dietetic Assoc. 1987, 87 (9), 11641171.
84. Jensen, H.; Madsen, J.L. Diet and Cancer. Acta. Med. Scand. 1988, 223,
293304.
85. Greenwald, P.; Lanza, E.; Eddy, G.A. Dietary Fiber in the Reduction of
Colon Cancer Risk. J. Am. Dietetic Assoc. 1987, 87, 11781179.
86. Rosen, M.; Nystrom, L.; Wall, S. Diet and Cancer Mortality in the Counties
of Sweden. Am. J. Epidem. 1988, 127, 4249.
87. Reddy, B.S.; Maeura, Y.; Wayman, M. Eect of Corn Bran and
Autohydrolized Lignin on 3,2 dimethyl-4-aminobiphenyl-induced Intestinal
Carcinogenesis in Male F344 Rats. J. Natl. Cancer Inst. 1983, 71, 419.
88. Sakagami, H.; Kohno, S.; Takeda, M.L.; Nakamura, K.; Nomoto, K.; Ueno, I.;
Kanegasaki, S.; Naoe, T.; Kawazoe, Y. O
2
Scavenging Activity of Lignins,
Tannins, and PSK. Anticancer Res. 1992, 12, 19952000.
89. Vikse, R.; Mjelva, B.B.; Klungsoyr, L. Reversible Binding of the Cooked
Food Mutagen MeIQx to Lignin-enriched Preparations from Wheat Bran.
Fd. Chem. Toxic. 1992, 30, 239246.
90. Asanoma, M.; Takshi, K.; Miyabe, M.; Yamamoto, K.; Yoshimi, N.; Mori,
H.; Kawazoe, Y. Inhibitory Eect of Topical Application of Polymerized
Ferulic Acid; a Synthetic Lignin; on Tumor Promotion in Mouse Skin, Two
Stage Tumorigenesis. Carcinogenesis 1994, 15, 20692071.
91. Powell, W.J.; Whittaker, H. J. Chem. Soc. 1924, 125, 357364.
92. Klason, P. Arkiv. Kem. Min. Geol. 1908, 3 (#5), 1.
93. Klason, P. Ber. 1923, 56 [B], 300.
94. Paschke, F. Cellulosechemie 1922, 3, 1921.
95. Paschke, F. Cellulosechemie 1923, 4, 3132.
96. Willsta tter, R.; Zeichmeister, L. Ber. 1913, 46, 2401.
97. Beckmann, E.; Liesche, O.; Lehmann, F.; Angew, Z. Chem. 1921, 34, 285.
98. Brauns, F.E. The Chemistry of Lignin, Chapter 10, Academic Press: New York,
1952.
99. Fisher; Homan, Z. Physiol. Chem. 1937, 245, 139.
MODIFICATION OF LIGNIN 285
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
100. Schmidt; Zeiser. Ber. 1931, 67, 2120; Schmidt; Zeiser; Dippold. Ber. 1937, 70,
2102.
101. Aulin-Erdtman, G.; Hegbom, L. Evensk Papperstid. 1958, 61, 187.
102. Adler, E.; Morton, J. Acta Chem. Scand. 1959, 13, 75.
103. Spencer, E.Y.; Wright, G.F. J. Am. Chem. Soc. 1941, 63 (#7), 20172020.
104. Pechmann; Burkard Ber. 1900, 33, 3594; Hansen. Ber. 1931, 64B, 943.
105. Kratzl, K.; Wittmann, E. Monatsh. 1954, 85, 7.
106. Adler, E.; Bjorkgvist, K.J.; Haggroth, S. Acta Chem. Scand. 1948, 2, 93.
107. Gierer, J.; Wallin, N.-H. Acta Chem. Scand. 1966, 20 (#8), 2059.
108. Brauns, D.E. J. Am. Chem. Soc. 1939, 61, 2120.
109. Brauns, D.E. Paper Trade J. 1940, 111, 35.
110. Urban, H. Cellulosechemie 1926, 7, 73.
111. Sarkanen, K.V.; Ludwig, C.H. Lignins, Occurrence, Formation, Structure, and
Reactions; John Wiley & Sons: New York, 1971; p. 519.
112. Carroll, J.H.; Wallin, H.C. Can. Patent 707,382, April 6, 1965.
113. Ball, F.J.; Doughty, J.B.; Vardell, W.G. U.S. Patent 3,185,654, May 25, 1965.
114. Morton, J.; Adler, E.; Marton, T.; Falkehag, S.I. Lignin Structure and
Reactions. Advances and Chemistry Series 1966, 59, 125.
115. Allan, G.G.; Halabisky, D.D. Pulp and Paper Magazine of Canada 1970, 71,
T50.
116. Robert Northey. Georgia Pacic Corporation. Personal Communication,
1997.
117. Fischer, F.; Schroder, H. Briennsto-Chem. 1930, 2, 2713.
118. Okabe, J.; Hachihama, Y. J. Chem. Soc. Japan, Ind. Chem. Sect. 1955, 58,
779; Chem. Abstr. 1956, 50, 10400.
119. Traynard, Ph; Robert, A. Bull. Soc. Shim. France 1952, 746.
120. Hibbert, H.; Marion, L. Can. J. Res. 1930, 3, 130.
121. Falkehag, S.I. U.S. Patent 3,697,497, 1972.
122. Gelfand, E.O.; Tushina, L.F. Arkhang. Lesotekn. Inst. 1973, 49. Balcere, D.
Khim. Drev. 1973, 72.
123. Nagaty, A.; Mansour, O.Y. Am. Dyestu Reptr. 1979, 68, 64.
124. Tramontini, M. Synthesis 1973, 12, 703775.
125. Wiest, E.G.; Baion, W.J. U.S. Patent 2,709,696 (to E. I. duPont de Nemours
Comp.), 1955.
126. Ball, J.C. Jr. U.S. Patent 2,863,780, 1958.
127. Shorygina, N.N.; Grushnikiv, O.P.; Tychima, V.D. Izvest. Akad. Nauk
S. S. S. R., Ser. Khim. 1967, 317; Chem. Abstr. 1967, 67, 3869.
128. Shorygina, N.N.; Sergeeva, L.L.; Lopatin, B.V. Izvest. Akad. Nauk S. S. S. R.,
Ser. Khim. 1967, 392; Chem. Abstr. 1967, 67, 3871d.
129. Bailey, C.W.; Dence, C.W. Tappi 1969, 52, 491.
130. Reeves, R.H.; Pearl, I.A. Tappi 1965, 48, 121.
131. Sarkanen, K.V.; Ludwig, C.H. Lignins; Occurrence, Formation, Structure,
and Reactions, ISBN 0-471-75422-6; J. Wiley, 1971; p. 821.
132. Rex, R.W. Nature 1960, 188, 1185.
133. Zeigerson, E.; Block, M.R. U.S. Patent 3,962,208, 1976.
134. Sarkar, P.B. J. Indian Chem. Soc. 1934, 11, 777.
135. Dence, C.W.; Meyer, J.A.; Unger, K.; Sadowski, J. Tappi 1965, 48, 148.
286 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
136. Burt Delacala, Westvaco Agricultural Chemicals Division, Charleston, South
Carolina, Personal Communication, 1997.
137. Norman, R.O.C.; Taylor, R. Chapter 3 In Electrophilic Substitution in
Benzenoid Compounds; Elsevier: London, 1965.
138. Nord, F.F.; de Stevens, G. Naturwissenschaften 1952, 39, 479.
139. Kurschner, K.; Peikert, H. Tech. Chem. Papier Zellsto-Fabr. 1934, 31, 1, 17,
53, 69, 73, 85; Chem. Abstr. 1935, 29, 5441.
140. Traynard, Ph.; Robert, A. Bull. Soc. Chim. France 1952, 746.
141. Friese, H.; Furst, H. Chem. Ber. 1937, 70, 1463.
142. Lieser, Th.; Schaack, W. Chem. Ber. 1950, 83, 72.
143. Konig, F. Cellulosechemie 1921, 2, 93, 105, 117.
144. Chuksanova, A.A.; Grusnikov, O.P.; Shorygina, N.N. Izvest.
Akad. Nauk S.S.S.R., Otdel Khim. Nauk 1961, 1810; Chem. Abstr. 1962,
56, 7549.
145. Kurschner, K. Zellsto-Faser 1935, 32, 87; Chem. Abstr. 1935, 29, 6755.
146. Shorygina, N.N.; Sergeeva, L.L.; Lopatin, B.V. Izvest. Akad. Nauk S.S.S.R.,
Ser. Khim 1967, 392; Chem. Abstr. 1967, 67, 3871d.
147. Kee, M.L. Ph.D. Dissertation, McGill University, Montreal, Canada, 1968.
148. Shorygina, N.N.; Mikhailov, N.P.; Lopatin, B.V. Khim. Prirod. Soedim.,
Akad. Nauk S.S.S.R., 1966, 2, 58; Chem. Abstr. 1935, 29, 6755.
149. Ref. missing
150. Sarkanen, K.V.; Ludwig, C.H. Lignins, Occurrence, Formation, Structure, and
Reactions; John Wiley & Sons: New York, 1971; pp. 251, 272, 487507, 535,
659, 818821.
151. Adler, E.; Marton, J. Acta Chem. Scand. 1959, 13, 75.
152. Gaslini, F. Tappi 1958, 41, 162A.
153. Kin, Z. Przeglad Papier 1960, 16 (5), 131.
154. Suter, C.M.; Bair, R.K.; Bordwell, F.G. J. Org. Chem. 1945, 10, 470.
155. Tilghman, B.C. Br. Patent 2924, 1866.
156. Pedersen, N. Papier-Ztg. 1890, 15, 422.
157. Hawley, L.F.; Wise, L.E. The Chemistry of Wood, A.C.S. Monograph Series,
The Chemical Catalog Co., Inc., New York, 1926.
158. Lindsey, J.B.; Tollens, B. Ann. 1892, 267, 341.
159. Holmberg, B. Svensk Kem. Tid. 1935, 57, 257.
160. Heden, S.; Holmberg, B. Svensk Kem. Tid. 1936, 58, 207.
161. Rogers, W.F. Compositions and Properties of Oil Well Drilling Fluids, 3rd Ed.;
Gulf Publishing Comp.: Houston, TX, 1963.
162. King, E.G.; Adolphson, C. U.S. Patents, 2,935,473 and 2,935,504, 1960.
163. Papadakis, M. La Revue Des Materiax de Construction, No. 519, December,
1958.
164. West Virginia Pulp and Paper Co. (Westvaco), Polychemicals Div., Technical
Bulletin 300.
165. West Virginia Pulp and Paper Co. (Westvaco), Polychemicals Div., Technical
Bulletin 102.
166. Halstead, W.J.; Chaiken, B. Public Roads 1961, 31, 126.
167. Sarkanen, K.V.; Ludwig, C.H. Lignins; Occurrence, Formation, Structure,
and Reactions, ISBN 0-471-75422-6; J. Wiley, 1971; pp. 849860.
MODIFICATION OF LIGNIN 287
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
168. Anon. Chem. Eng. News 1960, 38 (26), 40; Banko, J. Tappi 1961, 44, 849.
169. Northey, R.A. Low Cost Uses of Lignin. In Engineering Technologies for
Materials and Chemicals from Biomass, Rowell, R.M., Ed.; ACS Symposium
Series 476, American Chemical Society, Washington, DC, 1992.
170. King, E.G.; Adolphson, C. U.S. Patent 3,109,742, 1963.
171. Sherman, W.A. Paper Trade J. October 1950, 12, 19.
172. Anthone, R.; Parks, M.P. U.S. Patent 4,001,003, 1977.
173. West Virginia Pulp and Paper Co. (Westvaco), Polychemicals Div., Technical
Bull. 306B.
174. Greenacher, C.; Matter, M. U.S. Patent 2,490,953, 1949; Geary, R.J. U.S.
Patent 2,858,250, 1958.
175. Bonewitz, P.W.; Fults, E.H.; Hockett, S.W. U.S. Patent 2,826,522, 1958.
176. Kaye, S. U.S. Patent 3,317,431, 1967.
177. Lin, S.Y. Progress in Biomass Conversion, Academic Press: Orlando, FL, 1983;
Vol. 4, p. 31.
178. Tormala, P.; Lindberg, J.J.; Koivu, L. Paperi ja Puu 1972, 159164.
179. Lindberg, J.J.; Bulla, I.; Tormala, P. J. Polym. Sci. Symp. Ser. 1975, 53,
167171.
180. Blount, D.H. U.S. Patent 4,051,151, 1977.
181. Freundlich, H. Colloid and Capillary Chemistry, English Translation,
Methuen, Ltd.: London, 1926.
182. Brinker, C.J.; Scherer, G.W. Sol-Gel Science, Academic Press, Inc.: Boston,
Mass. 1990.
183. Doughty, J.B. U.S. Patent 3,081,293 (to Westvaco), March 12, 1963.
184. March, J. Advanced Organic Chemistry, 4th Ed.; J. Wiley and Sons, New
York, 1992b; p. 659.
185. Allan, G.G. Unpublished work.
186. Kreitsberg, Z.N.; Sergeeva, V.N.; Grabovskii, Y.K. Khim. Perarabotka i
Zashchita Drevesiny, Riga, Akad. Nauk. Latv. SSR, 1964, 81.
187. Nifanter, E.E.; Fursenko, I.V. U.S.S.R. Patent 181,108, Jan. 13, 1965.
188. Paschke, F. Cellulosechemie 1992, 3, 19.
189. Tropsch, H. Gos. Abh. Kennt. Kohl. 1923, 6, 301.
190. Tronov, B.V.; Pershina, L.A.; Morozova, V.M.; Kovapenko, A.V.; Galochkin,
A.I. Gidroliz. i Lesokhim. Prom. 1961, 14, 5, 10.
191. Kashima, K.; Oiwa, K. Yuki Gosei Kagaku Kyokai Shi 1959, 17, 221, 441.
192. Tronov, B.V.; Pershina, L.A.; Galochkin, A.T. U.S.S.R. Patent 164,276,
Dec. 23, 1963.
193. David, N.S.; Hon, Ed.; Graft Copolymerization of Lignocellulosic Fibers,
ACS Symposium Series #187, Am. Chem. Soc. 1982, ISSN 0097-1656; 187.
194. Chem. Eng. News 1984, 62 (#39), 1920.
195. Koshijima, T.; Muraki, E. Zairy, O. 1967, 16 (#169), 834838.
196. Phillips, R.B.; Brown, W.; Stannett, V.T. J. Appl. Polym. Sci. 1971, 15,
29292940.
197. Koshjima, T.; Muraki, E.; Nihon Mokuzai Gakkaishi 1964, 10, 44 (#3),
110115.
198. Koshjima, T.; Muraki, E.; Nihon Mokuzai Gakkaishi 1964, 10 (#3), 116119.
288 MEISTER
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
199. Simionescu, Cr.; Ceratescu-Asandei, A. Stoleru, Cellulose. Chem. Tech. 1975,
9 (#4), 363380.
200. Lawton, E.J.; Bellamy, W.D.; Hungate, R.E.; Bryant, M.P.; Hall, E. Science
1951, 113, 380.
201. Hachihama, Y.; Jyodai, S. Mem. Inst. Sci. and Ind. Res. Osaka Univ. 1948, 6,
74.
202. Lenz, R.W. Organic Chemistry of Synthetic High Polymers, ISBN 470-52630-0,
Interscience, 1967, pp. 161172.
203. Meister, J.J.; Patil, D.R. Macromolecules 1985, 15591564.
204. Meister, J.J.; Patil, D.R.; Channell, H. Ind. Eng. Chem. Prod. Res. Dev. 1985,
24 (#2), 306313.
205. Meister, J.J.; Patil, D.R.; Augustin, C.; Lai, J.Z. In Lignin, Properties
and Uses, Simo Sarkanen, Wolfgang Glasser, Eds.; American Chemical
Society, 1989; American Chemical Society Symposium Series, Vol. 397.
206. Meister, J.J.; Li, C.T.; Tewari, K.K.; Simoliunas, S. Spring National Meeting
of the American Chemical Society, April 2227, 1990 at the Symposium on the
Chemical Modication of Biopolymers, Boston, MA. American Chemical
Society Abstracts: April 1990, 199, p. 418.
207. Meister, J.J.; Chen, M.J. Macromolecules 1991, 24 (#26), 68436848.
208. Gunnells, D.W.; Gardner, D.J.; Chen, M.J.; Meister, J.J. Proceedings of the
American Chemical Society, Division of Polymeric Materials: Science and
Engineering, 1992, 67, 227. ACS Abstracts: August 1992, 204, p. 125.
Presented at the Fall Meeting, 1992, Washington, D.C.
209. Milstein, O.; Gersonde, R.; Huttermann, A.; Chen, M.J.; Meister, J.J. Appl.
Environ. Microbio. 1992, 58 (#10), 32253232.
210. Chen, M.J.; Meister, J.J.; Gunnells, D.W.; Gardner, D.J. Process for Coupling
Wood to Thermoplastic Using Graft Copolymers. Adv. Polym. Technol.
1995, 11 (2), 97109.
211. Advanced Technology Applications to Eastern Hardwood Utilization, 1992
Progress Report #5 to the U.S. Department of Agriculture, pp. 6674,
Department of Forestry, Michigan State University Press.
212. Meister, J.J.; Chen, M.-J.; Gunnells, D.W.; Gardner, D.J.; Milstein, O.;
Gersonde, R.; Hutterman, A. Graft copolymers of lignin with 1-ethenylben-
zene. II. Properties, Macromolecules 1996, 29 (5), 13891398.
213. Meister, J.J.; Aranha, A.; Wang, A. Proceedings of the American Chemical
Society, Division of Polymeric Materials: Science and Engineering, Chicago,
August 2326, 1993.
214. Meister, J.J.; Chen, M.J.; Lou, Y.; Aranha, A.; Zhao, Z. Bio/Environmentally
Degradable Materials Society, Second International Meeting, Chicago,
August 1920, 1993.
215. Demaret, V.; Glasser, W.G. Polymer 1989, 30 (#3), 570575.
216. DeOliveria, W.; Glasser, W. Polym. Prepr. 1990, 31 (#1), 655656.
217. Kelley, S.S. Incorporation of Lignin Copolymers into Polyurethane Materials,
Ph.D. Thesis, Virginia Polytech. Inst. State Univ. 1987, Univ. Microlms Int.
#DA8814587.
MODIFICATION OF LIGNIN 289
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1
D
o
w
n
l
o
a
d
e
d

A
t
:

2
0
:
1
9

2
6

J
a
n
u
a
r
y

2
0
1
1

You might also like