Sunteți pe pagina 1din 50

Wood Sci. Technol.

1 1 : 1 6 9 - 2 1 8 (1977)

Wood Science

anci Technology
9 by Springer-Verlag 1977

Academy Lecture
presented at the
Fifth Plenary Meeting of the International Academy of Wood Science,
Copenhagen, Denmark, June 19, 1976

Lignin C h e m i s t r y - P a s t , P r e s e n t a n d F u t u r e
Erich Adler
Department of Organic Chemistry, Chalmers University of Technology and University of
G6teborg, G6teborg, Sweden

Summary. Some pertinent results and views from the earher history of lig'nin chemistry, pointing to the importance of the arylpropane skeleton, are outlined. Later development, beginning
with the dehydrogenation theory and experimental studies on the dehydrogenative polymerization of p-hydroxycinnamyl alcohols, is then reviewed. Finally, recent degradative work resulting
in a detailed picture of lignin structure is discussed.

Introduction
As stated in the regulations of the Academy, the Academy Lecture shall "consist of
a review of what has transpired in the subject matter field over the years, what is
currently being done and future needs." Lignin chemistry in its broadest sense includes a number of different aspects such as the formation of the polymer and its
distribution in the wood, its structural chemistry and its reactions, its physical
properties, its behavior in pulping and bleaching processes and the problem of its
ut12ization. In this review, I wish to concentrate upon the structural chemistry of
lignin, not only because the efforts made in our laboratory over a number of years
were largely devoted to this aspect, but also because it is fundamental to most of
the other aspects mentioned.
It will not be possible to cover all important contributions to this subject and I
therefore want to refer to two recent publications in the area, namely the monograph "Constitution and Biosynthesis of Lignin" by IC Freudenberg and A. C. Neish
[1968] and the comprehensive book "Lignins" edited by K. V. Sarkanen and
C. I-L Ludwig [1971].

Some data from the earlier history of I'~nln chemistry


Anselme Payen [1838] was fh-st to recognize the composite nature of wood. He
found that treatment of wood with nitric acid removed part of the wood substance
and left behind a fibrous material which he named "cellulose". He also noted that

170

E. Adler

the carbon content of wood and, consequently, that of the removed material was
higher than that of the cellulose which had the same composition as starch. He
spoke of the carbon-rich substance as an "encrusting material" which embedded the
cellulose in the wood; the term '~lignin" was introduced by F. Schulze [1865]
(Fig. 1). E. Erdmann [1868] observed that catechol and protocatechuic acid were
formed on alkali fusion of wood and concluded that the non-cellulosic constituent
was aromatic. In 1890, Benedikt and Bamberger demonstrated that methoxyl
groups were present in woody tissues but were lacking in pure cellulose.
A. Payen (1838)

"encrusting material"

F. Schutze (1865]

"[ignin"

E. Erdmann(1868)

atkati fusion of wood produces


COOH
~OH

nd

OH

[~OH
OH

E Bamberger (1890)

tignincontains OCH3

P. Ktason (1897)

[ignin is structura[ly retated to


CH20H
I

HC
1
CH

OH

Conif~ry[~ c ehol
. Fig.

HC=O
I

HC

II

CH

OH

Coniferaldehyde

1.

Tiemann and Mendelsohn [1875] suggested the possibility that coniferin, a glucoside of coniferyl alcohol (Fig. 1), might be a transformation product of the "aromatic atom complex" present in wood. The development of new pulping processes,
especially the sulfite process, strongly stimulated interest in the reactions involved,
and in connection with his studies on the composition of the lignin sulfonates,
P. Klason [1897] advanced the idea that lignin was chemically related to coniferyl
alcohol. He found support for this view in his observation that heating of coniferyl
alcohol with acidic bisulfite solutions produced a sulfonic acid which he believed
was similar to lignin sulfonate. Remarkably enough, as early as 1907 Klason considered lignin to be a high-molecular substance and ten years later he proposed that
the coniferyl alcohol units might be linked together by a continuous condensation
between the alcoholic and phenolic hydroxyl groups. He also assumed that part of
the lignin was made up of units of coniferaldehyde or of an hydroxyconiferyl alcohol
rather than coniferyl alcohol. Although Klason was unable to produce solid experi-

Ligninchemistry-past, present and future

171

mental evidence for his structural views, the basic coniferyl alcohol idea undoubtedly
has greatly influenced the mind of later hgnin chemists.
Isolation of litmin
According to Klason [1908], removal of the carbohydrates from wood by hydrolysis with strong sulfuric acid yields a brown-colored lignin product as an insoluble
residue. Carbohydrate hydrolysis under milder conditions was achieved by Willstutter and Zechmeister [1913] who used cold 40 % hydrochloric acid, and by Urban
[1926] who worked in Freudenbergs laboratory and applied a 3 : 1 (v/v) mixture of
36 % hydrochloric acid and 80 % phosphoric acid. A "cuproxam lignin" was prepared [Freudenberg et al. 1928] by alternating treatments of wood with boiling 2 %
sulfuric acid and cuprammonium hydroxide.
In these and similar preparations the use of strong acids could be expected to
cause chemical changes such as condensation reactions. Bj6rkman's method of
extracting lignin with neutral solvents and without using elevated temperatures
therefore marked a major advance [Bj6rkman 1956, 1957a, b; Bj6rkman, Persson
1957a, b]. The method is based on the finding that about 30% of the lignin becomes extractable with dioxane-water (9 : 1), if wood, suspended in toluene, is
finely disintegrated in a vibratory ball mill. The "Bj6rkman lignin" or "milled
wood lignin" (MWL), a pale tan powder, still contains a few percent of carbohydrates, but modifications of the original purification procedure have been described which afford spruce lignin preparations containing less than 0.2 % [Harkin,
J. M., quoted in Freudenberg, Neish 1968, p. 52] and as little as 0.05 % [Lundquist,
Simonson 1975; Lundquist et al. 1977] of carbohydrates.
Pew [1957; Pew, Weyna 1962] showed that wood which has been subjected to
5-8 hours' dry milling in a vibratory ball mill, can be freed from most of its polysaccharides by treatment with a commercial glycosidase. The remaining lignin still
contained 12-14 % of carbohydrates, however. Chang et al. [1975] treated wood
meal which had been milled under toluene in a vibratory ball mill for 48 hours with
an enzyme preparation possessing high cellulolytic and hemicellulolytic activities
and isolated two fractions of "cellulolytic enzyme lignin" by extracting the digested
material successively with 96 % and 50 % (v/v) aqueous dioxane. The yields of these
hgnin fractions obtained from spruce wood were 27.8 and 29.2 % of the Klason
hgnin of the wood, and their carbohydrate contents were 4.3 and 8.9 %, respectively.
For comparison, Bj6rkman lignin, prepared from the milled wood without any
enzyme digestion was obtained in a yield of 16.8 % and had a carbohydrate content
of 4.1%. Similarly, the yield of lignin isolated from a hardwood (sweetgum) by the
enzymatic method was about four times that of the Bj6rkman lignin.
The milled wood lignins and the cellulolytic enzyme lignins are the best lignin
preparations now available. In fact, oxidative degradation of methylated Bj6rkman
lignin from spruce and of methylated spruce wood gave the same aromatic carboxylic acids in closely similar yields [Erickson, Larsson, Miksche 1973b]. In this sense,

172

E. Adler

milled wood lignins can be regarded as being representative of the lignin in the wood.
However, they are not identical with the lignin in situ. Certainly, fragmentation of
the lignin macromolecules is brought about by the vibratory ball milling and, therefore, the average molecular weights found for milled wood lignins, Mw = 15 000 for
a spruce MWL and 16000 for a sweetgum MWL [Chang et al. 1975], will be much
lower than those of the original lignins. Furthermore, certain chemical changes,
such as an increase in the number of free phenolic hydroxyl and a-carbonyl groups,
are taking place during the milling [Pew 1957; Chang et al. 1975].
The arylpropane skeleton
During the period of 1926-1932 a number of structural hypotheses were published
by IC Freudenberg. They were based mainly on the elemental compositions of the
spruce lignin preparations available at that time and a number of their reactions.
The building unit seemed to contain a gualacyl nucleus substituted with an oxygenated C 3 sidechain, in agreement with Klason's basic idea_ In addition to the oxygen
of the methoxyl group, there were two further oxygen atoms, one of which belonged to a hydroxyl group, whereas the second apparently was an ether oxygen.
Freudenberg was looking for a simple architecture, similar to that found earlier in
other natural macromolecules, with one or a few closely related building stones
linked together by continuous condensation. Among a number of different proposals those indicated in Fig. 2 are of special interest.
H2~ /

H2 c /

H COH

H C--O

HCOH

HCOH

'y

CH20H
/

13 H~OH
I
HCOH

.....

>

H~OH

HC--O
I
HCOH

HCOH

OH
1

OH

OH
3

HOC /
I
HCOH

~=o
HCOH

OH
,~

F ~ 2.

t
...... >

HOC ~

HC--O

oH

HCOH

OH

.....

_,.

HC-~O

OH
5

Lignin chemistry-past, present and future

173

According to these proposals [Freudenberg, ~


1932], the monomer, regarded
to be guaiacylglycerol (2.1) or its equivalent 2.4, is converted by C~C condensation
into a polymer (2.2 and 2.5), and the latter reacts further by loss of water to give
the cyclic phenyl ether structures shown in formulae 2.3 and 2.6.

Permanganate degradation
Freudenberg et al. [1936, 1938] heated spruce li~nin or spruce wood with 70%
aqueous potassium hydroxide in order to bring about hydrolytic cleavage of ether
linkages and subsequently protected the phenolic groups liberated by methylation.
Permanganate oxidation of the methylated product at pH 6 - 7 gave veratric acid in
a yield of about 8 % of the lignin and minor amounts of isohemipinic and dehydrodiveratric acids (Fig. 3). The formation of isohemipinic acid seemed to support the
occurrence of 3'-5 or ~-5 condensed structures as depicted in schemes 2.3 and 2.6,
but the comparatively high yield of veratric acid indicated that noncyclic ether
bridges between a sidechain hydroxyl and the phenolic hydroxyl group of the adjacent unit were also important. The 3,4-dimethoxybenzoic acid structure of the
degradation products was in harmony with the guaiacylpropane concept.
Permanganate oxidation

COOH

z KOH,;ZO"

Spruce L ignin

3. KMnQ~

p.

OCH3
OCH3
ver~tr.: a c i d
COOH

Fig. 3.

COOH

HOOC

OCH3
OCH3
isoh~mipinicacid

COOH

C H 3 0 ~ O C H 3
OCH3
OCH3
debydrc-diverattic acid

Nitrobenzene oxidation
Similar information was gained from the oxidative degradation of lignin with nitrobenzene in the presence of hot alkali. Spruce wood gave about 25 % of vanillin,
based on the Klason lignin content of the wood [Freudenberg et al. 1940], whereas
mixtures of vanillin and syringaldehyde were obtained from hardwoods [Creighton
et al. 1944]. In addition to these two aldehydes, grasses afforded p-hydroxybenzaldehyde [Creighton, Hibbert 1944]. Later, small amounts of the latter aldehyde
were also found in the oxidation mixtures from softwoods and hardwoods, and
traces of syringaldehyde were shown to be formed from softwoods [Bland et al.
1950; Leopold, Malmstr6m 1952].

174

E. Adler

Nitrobenzene oxidation

H~O

H~.c//O

H~c~O

"~ "OCH3 CH30~OCH3


OH
OH
vg~ll~
syringoldehyde

OH
p-hydroxybenz~[dehyd
e

Fig. 4.

Hyctrogenolysis
The first isolation of degradation products which contained the complete C6C3
skeleton was reported by Harris and Adkins [1938]. These authors obtained fair
yields of propylcyclohexane derivatives (Fig. 5) when they subjected a methanol
lignin from aspen to catalytic hydrogenation under vigorous conditions.
Guaiacylpropane and syringylpropane derivatives were later obtained by Hibbert,
Pepper, Schuerch and other workers [cf. Hrutfiord 1971] on hydrogenolysis using
a variety of conditions.

Hydrogenotysis

Methanol

[ignin

I~

CH2_
~"

Fig. 5.

on (ell3)

(OH)

OH

Ethanolysis
The "I-Iibbert ketones" 6.1-6.4 [Cramer, Hunter, Hibbert 1939; West, Maclnnes,
Hibbert 1943] which arise when coniferous wood is refluxed with 2 % ethanolic
hydrochloric acid were the first degradation products with a guaiacylpropane structure. Although the yields of these compounds were less than 10 % of the ligrdn,
Ethonolysis

S p r u c e Wood

Fig. 6.

2 % HCt
in EtOH
retCux

I~

OH1

HCOC2H
5
I
C=O

C=O
,
CH2

CH3
C=O
I
C~O

OH2

OH3

01-t

Lignin chemistry-past, present and future

175

their isolation strengthened the view of the C6C 3 nature of the ii~min building stones.
Larger amounts of the Hibbert ketones, including their syringyl analogs, were obtained from hardwoods [West et al. 1943; Kulka et al. 1944].
The sidechain structures of these ketones certainly had to be regarded as modifications of the original structures caused by the acid treatment. This was obvious
already from the fact that C--CH 3 groups were present in the lignin of the wood in
only negli~ble amounts. Thus, the proper nature of the C3 sddechains and the
nature of the linkages between the C6C 3 units in ligrlJn were still open questions.
The dehydrogenation theory
A new development in lignin chemistry was initiated by H. Erdtman [1933a, b] who
was studying the oxidative dimerization of phenols, especially from the point of
view of the biogenesis of certain natural products.
It was known from earlier work that oxidants like ferric chloride m a y remove a
hydrogen atom from a phenol, thus producing a free radical which stabilizes by
radical coupling to give a biphenyl or a diphenyl ether. Erdtman expected that a
phenol carrying in the ortho or para position a sidechain with a double bond conjugated with the ring, would be able to undergo coapling at the ~-carbon atom. On
dehydrogenation of a phenol of this type, isoeugenol (7.1), with ferric chloride or
with a mushroom oxidase, Cousin and H6rissey [1908] had obtained dehydrodiisoeugenol, to which a biphenyl structure was attributed [H6rissey, Doby 1909]. Erdtman [1933 a, b, c] disproved the latter structure and arrived at the phenylcoumaran
structure 7.6 for the dehydrogenation product. He interpreted its formation as
indicated in Fig. 7 (which differs from the presentation given by Erdtman only by

~ H3
H~

HC 13

II

CH

~OCH
OH
I

CH

phenol ox/da.se ~

.~

HC"

CH

,~

3
03

0' 2

O~
"v-

C H--CHq2H3

CH-CH-CH 3

"T -oc

OH
Fig. 7.

oc.
HCI

HC - - 0

OH

0
g

176

E. Adler

the use of the resonance structures 7.2-Z4 for the primanqy formed radical).
Coupling of one radical, reacting as 7.4, with a second one, reacting as 7.3, results
in the formation of a quinonemethide intermediate (7.5) and involves the coupling
reaction at C# expected by Erdtman. Ring closure by nucleophilic attack of the
phenolic oxygen atom upon the methide carbon atom terminates the reaction. The
correctness of formula 7. 6 was confirmed by Freudenberg and Richtzenhain [1942].
Considering the fact that Freudenberg had discussed the presence in lignin of
phenylcoumaran structures (2.3, 2.6) similar to 7.6, Erdtman [1933a, b] advanced
the idea that lignin is formed by dehydrogenation of phenolic, a,/~-unsaturated
C6C3 progenitors of the coniferyl alcohol type.

The dehydrogenative polymerization of p-hydroxycinnamyl alcohols


Towards the end of the 1930's the oxidation of coniferyl alcohol with ferric chloride was investigated in Freudenberg's laboratory [Freudenberg 1939]. It was found
to result in an amorphous product which on treatment with hot alkali followed by
methylation and permanganate oxidation gave veratric acid and isohemipinic acid
(cf. Fig. 3). The same degradation acids were obtained from dehydrodiisoeugenol
(7.6) and from lignin. This suggested the possibility that lignin is formed by dehydrogenation of coniferyl alcohol and related compounds, in accordance with
Erdtman's hypothesis.
In consideration of the results obtained on nitrobenzene oxidation of lignin
(Fig. 4), the natural precursors converted into lignin by dehydrogenation could be
assumed to be the three p-hydroxycinnamyl alcohols shown in Fig. 8.
Lignin Precursors

CH

p-Coun'~l
alcohol

CH

Coniferyt
at~hol

Sinapyl
alcohol

Fig. 8.

Dehydrogenation of coniferyl alcohol with air in the presence of a mushroom


oxidase [Freudenberg, Richtzenhain 1943] or with laccase shown to be present in
the crude mushroom enzyme [Freudenberg et al. 1958; Higuchi 1958], or with peroxidase and hydrogen peroxide [Freudenberg et al. 1963] gave a "dehydrogenation
polymer" (DHP) which showed certain similarities with spruce Bj6rklnan hgnin. If
the reaction was interrupted before the polymer began to precipitate, a mixture of
dimeric, tfirnerie and oligomeric products was obtained. It was possible to isolate
about 30 different compounds and to determine their structures. Fig. 9 shows a
selection of three dimers ("dilignols") and a tetramer ("tetrnliEnol").

LiEnin chemistry-past, present and future


Dilignols

and

177

o[igolignots

OH

Hg=CH-C1420H
CH=CH-CH20H

OCH3

H2~OH

HC

CH

HCOH

[~OCH3
OH

OH

OH
3

H2~OH
HC
I
H
[ ~

~H20H

CH20H

HC

O--CH

.~
CH30

H2COH
H~--O
I
CH

C
OCH3

~OCH3
OH

Fig. 9.

The major products were the dilignols guaiacylglycerol-~-coniferyl ether (9.1),


dehydrodiconiferyl alcohol (9.2), which is quite analogous to compound Z6 derived
from isoeugenol, and D, L-pinoresinol (9.3) [Freudenberg, Schliiter 1955; Freudenberg 1966].
As indicated by Fig. 9, many of the dilignols and oligolignols, which could be
isolated from the model dehydrogenation system, contained unsaturated sidechains.
It could be assumed that lignin formation involved further dehydrogenation of these
lignols followed by coupling of the arising radicals with the formation of biphenyl
and diphenyl ether linkages. If Lignin actually were formed according to this principle, it would be expected to contain a large number of unsaturated side chains.
However, the sum of terminal units of the coniferyl alcohol type (!0.1) and the
coniferaldehyde type (10.2) in Bj6rkman lignin from spruce is rather limited.
End groups

CH20H
H

9 COCH 3
.....O
I
F~. 10.

H...~//O
CH
n
HC

"" ~ O C H 3
0
2

178

E. Adler

The presence of coniferyl alcohol groups in extracted spruce wood sections was
demonstrated by Lindgren and Mikawa [ 1957] by means of a specific color reaction.
Marton and Adler [1961] found a value of 6 - 7 for the total of units 10.1 and 10.2
by measuring their contribution to the UV spectrum of the lignin.
Since the well-known color reactions of wood with phlorogtucinol-HC1 or with
aromatic amines had been found to be specific for coniferaldehyde groups [Adler
et al. 1948a, b; Kratzl 1948], they were used for the quantitative estimation of
these groups. The stable yellow color produced by the reaction with p-aminobenzoic acid was preferred, and the application of this reaction to spruce Bj6rkman
lignin indicated the presence of 3 4 coniferaldehyde end groups (10.2) per 100 C9
units [Adler, Gierer 1957; Adler, Marton 1961; Adler et al. 1966]. The same figures
were derived from UV and IR spectroscopic studies [Adler, Marton 1959; Marton,
Adler, Persson 1961]; very recently they were confirmed by 1H NMR measurements
[Lundquist, Olsson 1977].
The limited number of unsaturated end groups (10.1 and 10.2) is of importance
for the understanding of the principles prevailing in the formation of hgnins by
dehydrogenative polymerization of coniferyl alcohol and the related cinnamyl alcohols shown in Fig. 8. The enzymatic dehydrogenation is a one-electron transfer
resulting in the formation of a resonance-stabilized phenoxy radical shown in Fig. 11.
Stabilization of the radical occurs by coupling to another radical in any of the positions of the unpaired electron given in resonance structures a-d. So far, however,
no products arising from resonance structure e have been encountered.
Dehydrogenation of conifery[ atcohol

CH2,0H
HC
it
CH
-(e',
~OCH
O-H

HC
II
CH

HC"
I
CH
0

3
a

0
b

OCH3

HC
II
CH

HC
II
CN

0
c

0
d

H~
H
QH
0

O
9

OCN3

Fig. 11.

The formation of a lignin molecule will begin with a dimerizafion of the radical
to give one of the dilignols shown in Fig. 9, several other dilignols also being possible (p. 176). The continued growth of the molecule will predominantly take place
by what has been called "end-wise" polymerization [Sarkanen 1971]. In the lignifying cell there will be a low stationary concentration of the monomer, for instance coniferyl alcohol. Therefore, dimerization of the monomer radicals will be
less favored than their cross-coupling with phenoxy radicals formed by dehydrogenation of the phenolic end groups of dili~ols or larger polymers. The process is
illustrated by two examples in Figs. 12 and 13. In Fig. 12, a coniferyl alcohol radi-

Lignin chemistry-past, present and future

179

cal in its resonance form b is attached by/3-0-4 coupling (for the nomenclature of
the coupling modes see Sarkanen and Ludwig [1971], p. 14-16) to an end group
radical a'. (The resonance hybride of phenoxy radicals of this type will include
forms a', c', d' and e', analogous to the corresponding forms of the monomer radical
shown in Fig. 11, but no/3-radical form, since the sidechains are saturated.) The
result of this coupling will be a quinonemethide (12.1). The latter will react further
by addition of a molecule of water to give a guaiacylglycerol-/3-aryl ether structure.
As will be shown later, this sequence of reactions is the most frequent one in the
biological synthesis of the lignm macromolecutes. The ~-0-4 coupling mode has been
predicted at an early stage by Erdtman [1950; Erdtman, Leopold 1948, 1949].
"End-wise~ polymerization

Y CH20H
p HC.
+
r
o~ CH
5

OCH3
.0

H~ H20H
,.

a'

OCH3

I
CH

OCH3

OCH3

.~fH n~--~_ C_C_C


, ~
~ ~k~_/~OCH3

I+H20

C-C-C

y -OCH3
OH
guoi 0 c ylgl ycer0l ~13- diaryl ether
3

C-C-C

OCH3
OH
gu o ia cylglycerol13- aryl ether
2

F~. 12.

Instead of water, a phenolic hydroxyl group may add to 12.1. In this case, a
guaiacylglycerol~,0-diaryl ether structure will arise. Benzyl (or a-) aryl ether groups
as indicated in this structure have been found by Freudenberg [1966] in oligolignols,
such as the tetralignol 9.4 formed on Ot vitro dehydrogenation of coniferyl alcohol.
Their formation constitutes an additive, nondehydrogenative mode of growth of the
lignin molecules.
In the second example (Fig. 13) a phenolic end group of the guaiacylglycerol-~aryl ether type (13.1) is dehydrogenated and to the arising radical 13.2 in its resonance form c' a coniferyl alcohol radical b is added. The 13-5 coupling product 13.3,

180

E. Adler

in its tautomeric form 13.4, undergoes intramolecular ring closure with the formation of a phenylcoumaran system (13.5).

%CH

~ Hc

H~OHoI~ocH3
I
H~OCH3
HCOH

- (H'~.,. e)

~H

CH3

~
3

HCO~-I
H'~OCH3
cI
2

H~H
~:OH.~OCH3
~~c~y~OCH3
H-OCH3

F~.13.

~OH [~J~OCH3

4 - - H~o~
H~OTgCH3

{~OCH3
r

H3

" ' H~H'~r/ "OCH3


[~OCH3
3

Continued dehydrogenative addition of monomer radicals to the phenolic end


groups by/3-0-4 or/]-5 coupling according to Figs. 12 and 13 would result in sequences of a linear shape. An additional type of end-wise monomer attachment is the
~1 coupling mode which will be discussed later (p. 190). Furthermore, 5,5-coupling
to give biphenyl structures and 4-0-5 coupling resulting in diaryl ether linkages are
important growth reactions (i9. 202), as they may give rise to branching. Biphenyl
and diaryl ether coupling probably take place preferentially between two end-group
radicals rather than between an end-group radical and a monomer radical. The
probability of either one of these two types of coupling reactions to occur, in
relationship to structure and oxidation potential of the reactants involved, has been
studied in model experiments [Erickson, Miksche 1972].
Branching of the polymer is also brought about by the additive formation of
benzyl aryl ether structures (cf. 9.4 and 12.3).

Tracer studies
The role of coniferyl alcohol and related compounds as precursors of lignin in plants
has been amply verified by experiments using the laC-labeUed substances. It was
shown that coniferin, the j3-glucoside of coniferyl alcohol, which was labelled in the
aglycone, when administered to a spruce plant, was converted into radioactive lignin
[Freudenberg et al. 1955; Kratzl et al. 1957]. Kratzl et al. [1957, 1959] demonstrated the formation of radioactive Hibbert ketones from lignin synthesized in
spruce which had been fed with coniferin, the aglycone of which was labelled with
14C in the a-position. Similarly, sinapic acid (Fig. 8), which was labelled at the

Lignir~chemistry-past, present and future

181

middle C-atom of the sidechain, was incorporated into the lignin of wheat, which
on ethanolysis yielded radioactive Hibbert ketones [Higuchi, Brown 1963]. A detailed discussion of this subject, which would be beyond the scope of this paper, is
found in the monograph by Freudenberg and Neish [1968].
Some analytical data
The extensive work of Freudenberg et al. on the in vitro dehydrogenation of coniferyl alcohol has provided a probable picture of reactions involved in lignin formation. Still, this work had the character of model experiments, and analytical and
degradative studies of lignin were necessary to confirm and to complement the
results of the "synthetic" approach.

Elemental analysis
Elemental analysis and methoxyl determination of Bj6rkman lignin from Norway
spruce indicated the following formula based on C9 [Freudenberg. In: Freudenberg,
Neish 1968, p. 113]:

C9HT.9202.40(OCH3)o. 92
Conksidering that the postulated starting material, coniferyl alcohol, had two oxygen atoms, excess oxygen could be written as water which had been added during
the polymerization (cf. Fig. 12):

C9H7.1202(H20)o.40(OCH3)0.92

(formula A)

The fact that the methoxyl content was less than 1.0 was ascribed to the copolymerization of coniferyl alcohol with minor amounts of the methoxyl-free p-coumaryl alcohol; in addition, a small proportion of sinapyl alcohol is incorporated.
Freudenberg [In: Freudenberg, Neish 1968, p. 81] assumed the ratio of these alcohols in the formation of spruce ljgnin to be 80 : 14 : 6, which would correspond to
0.80 + 2.0.6 = 0.92 OCH 3. The hydrogen content for such a mixture would be
10 - 0.92 = 9.08, and its composition:

C9H9.o802(OCH3)o.92

(formula B)

Subtraction B - A indicates that the formation of spruce lignin involved the loss
of 1.96 atoms of hydrogen and the addition of 0.40 molecules of water. These
characteristics are in harmony with the view that lignin is formed by dehydrogenative polymerization of a mixture of the three p-hydroxycinnamyl alcohols and that
a certain amount of water is added during the process.
It should be noted that on the basis of results obtained on oxidative degradation
(see p. 198), the ratio of guaiacyl : p-hydroxyphenyl : syringyl units in spruce lignin
has been estimated as 94 : 5 : 1 [Erickson et al. 1973 b] rather than that assumed by
Freudenberg (80 : 14 : 6). Indeed, methoxyl values of 0.95 and 0.96 per C9 have

182

E. Adler

recently been found for Bj6rkman lignin and "ceHulolyfic enzyme lignin" from
spruce [Chang et al. 1975].
For a Bj6rkman lignin from beech (Fagus silvatica) Freudenberg [In: Freudenberg, Neish 1968, p. 113] reports the composition C9H6.a302(H20)o.s3(OCH3)l.39
corresponding to a loss of 2.18 atoms of hydrogen, and for a grass lignin (alfalfa,
Medicago sativum) C9H7.2202(H20)o.41(OCH3)o.84 indicating a loss of 1.90 H.

Phenolic hydroxyl groups


Early analytical data had favored the view that the phenolic groups of the lignin
units were largely etherified (cf. p. 172 and Fig. 2), and the dehydrogenation studies
discussed above (p. 175) suggested the occurrence of/3-aryl ether structures (12.2), as
well as of noncyclic (12.3) and cyclic a-awl ether structures (13.5). Determination
of the amount of free phenolic hydroxyl groups should give a measure of the number of ether groups present.
Aulin-Erdtman [1954] made use of the differences in UV absorption of phenols
in neutral and alkaline solutions. The "ionization A e" method is useful in the investigation of conifer lignins and differentiates between various types of phenolic
structures such as those present in units with or without a keto group in the
a-position of the side-chain. The interpretation of A e curves of hardwood lignins
seems to be difficult.
Titration of phenolic hydroxyl in nonaqueous media has been reported by several
authors. Freudenberg et al. [1964] found 34 phenolic OH per 100C 9 units in spruce
Bj6rkman lignin, but this figure includes 3 - 6 carboxyl groups [Adler et al. 1968].
A specific method for the determination of phenolic groups in lignin is based on
the finding that phenols carrying a methoxyl group in the o-position are oxidized by
periodate (Fig. 14) with the formation of the corresponding o-quinone and one
molecule of methanol [Adler, Hernestam 1955; Adler et al. 1958]. The methanol is
conveniently determined by gas chromatography [Gierer et al. 1964; Johansson,
Miksche 1972]. The values found experimentally have to be corrected for the pre-

Periodotemethod
CH30..~OCH 3
O~

N~IO~

~CH30..~ O
O

FiB. 14.

sence of p-hydroxyphenyl units in the lignin which, of course, do not respond to


this method.
Application of the periodate method to samples of Bj6rkman lignin gave a corrected value of 25 phenolic hydroxyl groups per 100 C9 units [Adler et al. 1968].
In later measurements, however, a spruce Bj6rkman lignin gave a corrected value of
20.5 phenolic OH/IOOC 9 units [Chang et al. 1975]. The discrepancy is most probably explained by the fact that the wood sample from which the lignin was ex-

Lignin chemistry-past, present and future

183

tracted was ground in a highly efficient mill in the former case, whereas a less
efficaceous mill was used in the latter. As already mentioned (i9. 172), it has been
shown that vibratory ball milling of isolated lignins causes the liberation of phenolic
hydroxyl groups, presumably by cleavage of benzyl aryl ether bonds.
It therefore seems justified to conclude that the amount of units with a free
phenolic hydroxyl group in unaltered spruce lignin is < 20 pel 100 C9 units. Hence,
in a great majority of the guaiacylpropane units the phenolic hydroxyl is etherified.
The sample of milled spruce wood which had afforded the above-mentioned
Bj6rkman lignin with 20.5 phenolic OH/IO0 C9 units was also used for the preparation of "cellulolytic enzyme lignin" (CEL). Extraction of the enzyme-treated wood
with 96 % aqueous dioxane gave a lignin preparation (CEL 96) which according to
the periodate method had approximately the same phenolic hydroxyl content as the
Bj6rkman lignin. By subsequent extraction with 50 % aqueous dioxane a second
lignin fraction (CEL 50) with only 15 phenolic OH/100 C9 units was obtained.
Application of Aulin-Erdtman's A e method to these lignm preparations gave phenol
values which were closely similar to those determined with periodate [Chang et at.
1975]. A Bj6rkman lignin and two fractions of ceUulolytic enzyme lignin (CEL 96
and CEL 50) prepared from the same hardwood sample (sweetgum) were shown by
the periodate method to contain 14.5, 13 and 9 phenolic OH/100C 9 units, respectively. This indicates that the syringyl units in hardwood lignins are very extensively
ethefified [Chang et al. 1957]. It is noteworthy that the two CEL 50 lignin fractions
with the lowest phenol values had higher molecular weights (I~w), viz. 35,000
(spruce) and 33,500 (sweetgum), than the CEL 96 fraction from spruce and the
Bj6rkman lignins (Mw = 15,000 - 16,000) and the CEL 96 fraction from sweetgum
(Mw = 24,000).

Benzyl alcohol and benzyl ether groups


In the middle of the 1930's, B. Holmberg [Berg, Holmberg 1935; Holmberg 1936]
made the important suggestion, on the basis of simple model experiments, that
characteristic reactions of iignin, such as its sulfonation in sulfite cooking, its alkylation by alcoholic hydrochloric acid and its reaction with thioglycolic acid, were
reactions of benzyl alcohol or benzyl ether groups, often referred to as '~dae reactive
groups" of lignin. Numerous later investigations lent support to this suggestion
[Lindgren 1953; Adler, Lindgren 1952; Adler, Gierer 1955; Adler 1961a, b]. It has
further been shown that these groups are also involved in the initial reactions of
lignin during soda and kraft cookings [Gierer 1970; Miksche ]972].
B~nzytolcof~ and benzylaryl ether groups
c

c
HC-OH

FiB. 15.

~OC
0
I

OCH3

-C-C
H3

T
0

OCH3
2

184

E. Adler

When spruce Bj6rkman ligrtin, from which carbonyl and ethylenic groups had
been eliminated by catalytic hydrogenation [Marton, Adler 1961], was treated with
cold methanolic hydrochloric acid, it acquired 0.42 new methoxyl groups per
IOOC9 units [Adler 1958, 1961 a]. By comparison with the behavior of model
compounds it was concluded that this figure reflected the sum of benzyl alcohol
groups (15.1) and noncyclic benzyl aryl ether groups of type 15.2 or benzyl alkyl
ether groups.
A method supposed to be specific for benzyl alcoholic groups was subsequently
developed. It was based on the finding that suitable model benzyl alcohols were
readily oxidized by 2,3-dichloro-5,6-dicyano-benzoquinone (DDQ) to give the corresponding a-carbonyl compounds [Becket, Adler 1961 ]. Treatment of diazomethanemethylated Bj6rkman lignin (spruce)with DDQ generated a-carbonyl groups in an
amount of 16 CO-groups/lO0 C9 units as determined by UV spectroscopy [Adler et
al. 1966]. A color reaction [Gierer 1954] indicated that untreated Bj6rkman li~in
contained 6 benzyl alcoholic units (15.1) carrying a free phenolic hydroxyl group;
hence, there should be 16 - 6 = 10 benzyl alcoholic groups with an etherified
phenolic hydroxyl.
It was found in model experiments [Johansson, Miksche 1972] that phenolic
benzyl aryl ethers (I5.2, O=OH) are cleaved by cold dilute alkali and that both
phenolic and non-phenolic benzyl aryl ethers (15.2, 0 =O-C) are selectively hydrolyzed by a mild acid treatment. The numbers of these two types of ether linkages
in lignln could therefore be determined by measuring the increase in phenolic OH
which took place when lignin was treated as mentioned. The experiments indicated
that spruce Bj6rkman lignin per 100C 9 units contained at most 2 benzyl aryl ether
groups (15.2) of the phenolic type and 5-7 of the nonphenolic type [Adler et al.
1968]. These data confirmed the similar results obtained earlier in a different
manner by Freudenberg et al. [1964].
The total of groups 15.1 and 15.2 thus estimated, namely 24 groups per 100C9
units, is far below the above-mentioned value of 42 groups based on the uptake of
methoxyl on treatment of carbonyl-free lignin with methanolic hydrochloric acid.
Possibly, this may indicate that lignin contains, in addition to the non-cyclic benzyl
aryl ether groups (15.2), non-cyclic benzyl alkyl ether structures formed by addition
of a benzyl alcoholic group or a terminal carbinol group to a quinonemethide intermediate [Adler et al. 1966; cf. also Glasser, Glasser, Nimz 1976]. However, no
direct experimental evidence is available for the occurrence of such types of ether
linkages (a-O-a or a-O-3'). Cyclic a-O-7 linkages, as present in pinoresinol (9.3),
do not interfere, since pinoresinol was found not to react with cold methanolic
hydrochloric acid. The same is true for the cyclic benzyl aryl ether grouping present in the phenylcoumaran system, since the dihydroderivatives of dehydrodiisoeugenol (Z6) and dehydrodiconiferyl alcohol (9.2) are also stable under the conditions
used in the methylation of lignin [Adler 1961a]. Further investigations to clarify
the nature and the amount of the "reactive groups" in lignin would be desirable.

Lignin chemistry-past, present and future

185

Carbonyl groups
The infrared spectrum of lignin indicates the presence of minor amounts of conjugated as well as non-conjugated carbonyl groups. Oximation of spruce Bj6rkman
lignin as well as a volumetric borohydride method gave a value of about 20 carbonyl
groups per 100C 9 units [Adler, Gierer 1957; Marton, Adler, Persson 1961]. Examination of the changes in UV absorption taking place on reduction of the lignin with
sodium borohydride [Adler, Marton 1959] showed that this total number of carbonyl
groups included, per 100 C9, about 1 coniferaldehyde and 1 aryl-a-ketone unit with
free phenolic hydroxyl groups and, in addition, about 3 coniferaldehyde and 5 - 6
aryl-a-ketone units in which the phenolic hydroxyl groups were etherified. The same
method applied to a cellulolytic enzyme li~'nin (spruce) gave a value of 8 a-keto
groups, which, however, has to be regarded as an upper limit [Chang et al. 1975].
Similar results regarding the amount of conjugated carbonyl groups were obtained
by UV-spectroscopic examination of the hydrogenation of Bj6rkman li~in over a
Pd/BaSO 4 catalyst [Adler, Marton 1961; Marton, Adler 1961 ].
Initially, the difference between the total number of carbonyl groups (20 per
100 C9 units) and that of the conjugated carbonyl structures, s e., about 10 per
100 C9 units, was attributed to non-conjugated keto groups in the ~position of the
sidechains [Adler, Marton 1959]. Since the presence of glyceraldehyde groups in
lignin had been demonstrated (19. 190), Sarkanen [In: Sarkanen, Ludwig 1971,
p. 226-227] noted that these structures might account for the non-conjugated
carbonyl groups.

Degradation studies
It was obvious from the dehydrogenation remits discussed in a previous section that
the arylpropane units in lignin were interconnected by a variety of different types
of linkages. The final elucidation of lignin structure therefore had to aim at providing direct experimental proof of the existence of the proposed dimer structures in
lignin and determining their frequencies. A wealth of information regarding these
problems has been adduced during the past two decades by the use of various degradation processes.

Synthetic model compounds


The interpretation of the chemical behavior of lignin, especially in degradation procedures, is greatly facilitated by studying the behavior of appropriate model substances under the conditions used. A group of very useful model compounds is
shown in Fig. 16.

186

E. Adler
Synthetic model compounds

CIH2OH
HCOH
I
HCOH

~OCH3
OH

[OC"4
!

f2]

CH2OH [ ~ O C H 3
HC,
0
HCOH
[~
CH30""
OCH3
OH

EOC.,]

3
f ~]

CH20H [ ~ O C H 3
HC,
0
HC-O
CH2CH2CH
3
~ " OCFI3
~ "OCH3
OH

[OCN~]

5
f62

Fig. 16.

Guaiacylglycerol (16.1) and veratrylglycerol (16. 2) were synthesized [Adler,


Bj6rkqvist 1951, 1953; Adler, Yllner 1953], because they seemed to represent a
basic monomer structure, the side-chains possessing the same degree of oxidation as
that of coniferaldehyde. Also, they contained the characteristic benzyl alcoholic
group. Indeed, in sulfonation as well as in ethanolysis, these compounds showed a
llgnin-like behavior [Adler, Yltner 1953, 1954]. Both the D, L-erythro and the
D, L-threo forms of 16.1 and 16.2 were prepared [Adler, Gustafsson 1963; Miksche
et al. 1966].
Considering the possible importance of the fl-O-4 coupling in the biosynthesis of
lignin [Erdtman, Leopold 1948, 1949; Erdtman 1950], the veratrylglycerol-fl-guaiacyl
ether (16.4) was synthesized [Adler, Lindgren, Saed6n 1952], followed by the preparation of the phenolic analog (16.3) [Adler, Efiksoo 1955]. Freudenberg et al..
[1954, 1955] reported the synthesis of guaiacylglycerol-~-coniferyl ether, which had
been found among the dimeric dehydrogenation products of coniferyl alcohol.
Numerous methods for the preparation of model compounds of type 16.3 and 16.4
have been published later; for references see Lundquist and Remmerth [1975].
/3-Aryl ethers of syringylglycerol (indicated in formula 16.3) have also been synthesized [Miksche 1973]. Finally, the ct, fi-diaryl ethers 16.5 and 16.6 also became
available [Johansson, Mik~he 1972]. Among other widely used lignin models the
dehydrodiconiferyl alcohol (9.2) and its dflaydro derivative [Freudenberg, Hiibner
1952] may be mentioned. These models played an important role in the study of
the acidolytic degradation of lignin, which is discussed in the following section.

Acidolysis
Treatment of guaiacylglycerol (16.1) with refluxing ethanolic hydrochloric acid
yielded the "I-Iibbert ketones" 6.2, 6.3 and 6.4 [Adler, Yllner 1954] and, similarly,
the veratryl analogs of 6.1 and 6.2 were obtained from veratrylglycerol (16.2)
[Adler, Yllner 1952]. Arylglycerol units, however, could not be the major source
of the Hibbert ketones formed on ethanolysis of wood, because periodate oxidation

187

Lignin chemistry-past, present and future

of lignin produced only small amounts of formaldehyde [Lindgren, Saed6n 1952;


Adler, Yllner 1954; lmndquist, Lundgren 1972]. Therefore, the ethanolysis of the
13-guaiacyl ether of veratrylglycerol (16.4) was examined and found to give the veratryl analogs of Hibbert ketones [Adler, Pepper, Eriksoo 1957].
Since it had been found that refluxing of wood with dioxane-water (9 : 1) containing the equivalent of 0.2 M HC1, results in the formation of an ether-soluble oil
in addition to a high-molecular lignin product [Pepper, Baylis, Adler 1959], this
treatment, which was termed "acidolysis", was subsequently applied instead of
ethanolysis both to model compounds and to isolated lignin.
On 4 hours' acidolysis of 16.3 the 13-ether linkage was cleaved, guaiacol being
released, and furthermore, co-hydroxyguaiacylacetone (17.4) could be isolated in a
yield of 53 %. By analogy with Hibbert's ethanolysis reactions, the ketol 17.4 was
slowly further converted, yielding the isomeric ketols 17.5 and 17.6 (total yield,
15%), as well as small amounts of ketones 17.7 and 17.8. Only 3.5% of unchanged
starting material remained. The acidolytic cleavage of the/3-ether linkage in 16.3 is
assumed to proceed via a benzylium ion and an enol aryl ether, which is susceptible
to acid hydrolysis [Adler 1961 a; Adler, Lundquist, Miksche 1966; Lundquist, Lundgren 1972]. The conversion of the primary hydrolysis product 17.4 into the ketones
17.5-17.8 has been studied in detail [Lundquist, Hedlund 1967].
Actdo~ysis of Arytglycero~-B-oryl Ether Structure
Monorr~ric Acidc~ysis Products

C,~H
~HC--O-~c-c-c
aHCOH

HC=O

OCH3

~H2

~CH
2

CHpH

CH 3

C=O,
CH2

C=O,

HCOH

H- C

3
3
I

C,H3
H~OH
C::O

C---O

CH3
C--O
'H2

[
~OCH 3
OH
7

OH
5

F~. 17.

CH20H

HC--O

HC=0

C::X3,

H~

H~

cN

OH
9

~OCH 3
OH
8

tO

l]

When Bj6rkman lignin (spruce) was subjected to the same acidolysis treatment
[Lundquist 1970], the low-molecular portion of the resulting mixture could be re-

188

E. Adler

solved by gel f'titration into fractions containing monomeric, dimeric and oligomeric
compounds, respectively. In the monomer fraction the same ketones (1Z4-17.8)
as those formed from model compound 16.3 were detected [cf. also Lundquist, Kirk
t971], the predominating ketol 1Z4 being obtained in yields of 5 - 6 % of the tignin.
In addition, the presence of small amounts of homovanillin (1Z2) and formaldehyde
(17.3) was demonstrated. Both aldehydes also formed on acidolysis of model compound 16.3. This side-reaction can be regarded as a reverse Prins reaction of the
benzylium ion intermediate [Lundquist, Ericsson 1970]. These results constitute
clear evidence for the presence of substructures of the guaiacylglycerol-f-aryl ether
type (17.1) in lignm; the phenolic hydroxyl group can be free or etherified in an
acid-hydrolyzable fashion. The monomer fraction further contained small amounts
of ketol 17. 9 which must originate from p-hydroxyphenylglycerol~-aryl ether structures, and of coniferaldehyde (17.10) and p-coumaraldehyde (17.11).
In a similar monomer fraction obtained from the acidolysis of Bj6rkman lignin
from birch, Lundquist [1973 a] detected, in addition to most of the compounds shown
in Fig. 17, a number of their syringyl analogs. Ketol 17.4 was obtained in a yield of
3 % and its syringyl analog in a yield of 5 %. The yields of the syringyl monomers
were higher than those of the guaiacyl monomers, although the syringyl/guaiacyl
ratio in birch is about 1 : 1. This is due to the fact that part of the guaiacyl units
are linked to an adjacent unit by 5-5 and/%5 C~C bonds, which cannot occur in
syringyl units.
Dimeric Acidoiysis Products

I
c.2o.

i
.c.2o.

OH

CH2

OCH3

HC
~H

[
7

~LOC H3
OH 3

~H2

H~

CH

H2~

],.-OCH3

OCH3

Hq:-CH3

-CH3 H3

q:=-O

~r~ OCH 3
OH
5

OCH3
)H~
OH

CH30~[/OCH3
CH30

OH

OCH3
H?

CHt

OH
F i g . 18.

OQ43

O
"OCH3
OH
6

"~/

189

Lignin chemistry-past, present and future

From the dimer fraction obtained from spruce lignin, acidolysis compounds
18.1-18. 7 were isolated, and their structures were established. With the exception
of the trace constituents 18.2 and 18. 7, the dimeric fraction of birch lignin gave the
same guaiacyl compounds but, in addition, the corresponding syringyl analogs (with
one or two syringyl nuclei) of 18.3-18.6 and, furthermore, the stereoisomeric compounds D,L-syringaresinol and D,L-episyringaresinol (18.8). The analog of 18. 7
with two syringyt nuclei has been isolated as one of the products obtained on
degradation of beech wood with thioacetic acid [Nimz 1974] (19. 196).
The phenylcoumarone 18.1 and the stilbene 18.2 originate from a lignin substructure of type 19.1. This is obvious from the fact that the dihydro derivative
under acidolysis conditions is converted to a phenylcoumarone and a stilbene which
differ from 18.1 and 18.2 only by carrying a propanol instead of a ketol sidechain
[Adler 1957; Adler et al. 1959; Adler, Lundquist 1963; Lundquist, Hedlund 1971].
In both cases, the phenylcoumarone is formed in much higher yield than the stilbene.
The former product has a characteristic and very strong UV absorption. This pern'fitted its quantitative estimation which indicated that about 10% of the C9 units in
spruce lignin are connected to an adjacent unit by an a-0-4 as well as a/3-5 linkage,
giving rise to a phenylcoumaran system as present in structure 19.1 [Adler, Lundquist 1963].
Acldo~ysl5 of Phenyicoumaran Substructure

Hoc,h [~'ocH~

"~H~

c~oH
~,:o

C,
H20H
c--o

CH2

CH2

~.

HOCH2 II

~-C"ocH~

"c"-C"ocH~

H~ , ~CH~

H~O

C--O
OCH3
1

OH OCH3
181

H3
182

Fig. 19.
The phenylcoumarone 18.1 was the first dimeric lignin degradation product showing two complete C6C 3 skeletons [Lundquist 1964]. Its structure was unambiguously proven by synthesis [Lundquist 1970]. The acidolytic convermon of a hydroxymethyl-substituted phenylcoumaran into a methyl-substituted phenylcoumarone is
readily explained by a sequence of ring opening, allylic rearrangement and recyclization [Adler 1957; Adler, Lundquist 1963].
The "dimeric" compounds 18. 3-18. 6 [Lundquist, Miksche 1965; Lundquist
1970] all exhibit only one sidechain per two guaiacyl residues. It was postulated
that these compounds could arise from a 1,2-diguaiacyl-l,3-propanediol substructure

190

E. Adler

(20.1) incorporated into lignin by acid-hydrolyzable linkages. In fact, acidolysis of


synthetic 20.1 gave the expected products; a plausible mechanism for their formation has been presented [Lundquist, Miksche 1965; Lundquist 1973b]. Compound
20.1 and related compounds carrying one or two syringyl rather than guaiacyl
nuclei were isolated by Nimz [1965, 1966a, b] on "mild hydrolysis" of spruce and
beech woods (p. 193), and 20.1 was also encountered among the low-molecular products of coniferyl alcohol dehydrogenation [Freudenberg, Nimz 1966].

Acidotysisof 1,2-Di~rytpropane-l,3-dio[Substructure
oc..

HCOH
~OCFI
OH 31
acidolysis
C H 3
o.

CH

[~N OCH3
182

C=O

o.

O~HOCN3
18,(

o--c,

OCH3

I-tC--~--~ OH

~OCH3
OH 185

cH3o/, cH3/~oc~
Ho

oH

H/C<-o
185

/H

H--C%0

Fig. 20

The biogenesis of the lignin substructure 20.1 can be visualized as shown in


Fig. 21. A ~-1 coupling between a coniferyl alcohol radical and the radical of a
p-hydroxybenzyl alcohol end group gives a cyclohexadienone derivative 21.1. The
latter undergoes a reverse vinytogous aldol reaction and addition of water to the
quinonemethide group to give 20. I and a glyceraldehyde-2-aryl ether end group
(21.2) [Lundquist et al. 1967].
The ~3-1 coupling depicted in Fig. 21 thus involves the detachment of a sidechain,
which appears as a glyceraldehyde-2-aryl ether structure disrupting the growth of a
polymer chain. The second product (20.1), however, offers two phenolic hydroxyl
groups from which new chains may start. Experimental evidence for the presence
of glyceraldehyde end groups of type 21.2 was provided by the detection of pyruvaldehyde (= methylglyoxal, 22.2) in the acidolysis mixtures from spruce and birch
Bj6rkman lignin [Lundquist et al. 1967]. The mechanism of the acidolytic formation of pyruvaldehyde is presented in Fig. 22.
Colorimetric determination of the aldehyde formed on acidolysis of spruce and
birch lis
as well as of model compounds of type 21.2, indicated that only about

191

Lignin c h e m i s t r y - p a s t , present and future


Biogenesis of 1,2-Diory[propone-l,3~dio[

Substructure

-C-C

HOC~

HC"

OCH3

NC

CH

HC--CH

c.,o o

oc.,

HCOH~

CI-~O

OCN3
1

d'

+.,oI
I
C-C-C

H2CO H/z_~OCH3

FIOCH2
[ ~OCH3
Hr
I

HCOH

FI-'C-~o

cH3o~
OH

1,2- diary[-propone1,3-dio[

Fig. 2L

glycero[dehyde 2- aryt ether

20.1

AcidotysLs of Gtyceratdehyde-2-ary[ Ether Substructure

C--C--C

C-C-C

H2OH

HC
I
C

./%0 21,2

OCH3
0

-H~

~H2 ~ ' ~ O C H
I

H / C%0

H3o+

CH3
C=O
+
I
H/C"~0

Fig. 22.

C--C-C

[
]~OCH3
OH

PyruveJd~ayde

2 % of the C9 units of lignin were bound to glyceraldehyde as shown by formula


21.2 [Bemdtson et al. 1974]. The results of recent 1H NMR studies on the aldehyde
groups present in spruce liguin point to a similar value for the amount of glyceraldehyde groups [Lundquist, Olsson 1977]. From these results, the amount of ~-1
[inked C9 units (20.1) might also be concluded to be about 2 %. This figure, however, appears very low in view of the fair yields of degradation products of the
13-1 type obtained from beech wood [Nimz 1974]. It has been proposed by Sarkanen
[In: Sarkanen, Ludwig 1971, p. 226-227] that the unconjugated carbonyl groups in

192

E. Adler

spruce Bj6rkman lignin, which amount to about 10 carbonyl groups per 100C 9 units,
might be regarded as being present in gtyceraldehyde groups of type 21.2. This
proposal, however, does not Fred support in the analytical results mentioned above.
Further investigations will be needed to clarify the problem of the frequency of substructures 20.1 and 21.2.
In Fig. 23, the substructures which have been disclosed by the acidolysis procedure are summarized. Of these structures, the arylglycerol-/3-aryl ether structure
Substructures Derived from AcidolysisResutts

C
c

o~

CH30

CH30
/0

.2C/0\0H2
I
,

H2CO"

CH

HC --

CH2

H3

H2COH

CH30

OCH3
/0

CH30,~.OCH3
0
/
3

.0

O\

CH3~ 10 ~

He
5

H/C~o

OCH3

0
,5

F~. :Z3.

(23.1) undoubtedly is the most abundant one (cf. p. 202). This is true both for
spruce and birch lignins. In the latter case, syringylglycerol~-aryl ether structures
are involved in addition to the guaiacyl analogs 23.1. As already mentioned (p. 189),
C9 units which are linked to an adjacent unit by forming a phenylcoumaran system
(23.2) occur in spruce ligniu in an amount of about 10%. Similarly linked syringyt
units have been detected by Nimz [1974] in beech wood, the total of guaiacyl and
syringyl units of this type constituting 6 % of all C9 units (cf. p. 207).
A similar frequency (5 %) was attributed by Nimz to the/~-~ linkage of the
syringaresinol structure 23.3 demonstrated to be present in beech wood lignin.
D, L-syringaresinol and its stereoisomer, D, L-episyringaresinol, both represented by
formula 18.8, have also been found among the acidolysis products from birch
[Lundquist 1973]. The corresponding guaiacyl compounds, D, L-pinoresinol (9.3)
and D,L-epipinoresinol, however, could not be found in spruce or birch acidolysis

/_.it~ninchemistry-past, present and future

193

mixtures, although in lignin acidolysis experiments an amount of 0.1% of added


pinoresinol could be readily detected. The amount of pinoresinol structures in
lignin which can be liberated by acidolysis thus seems to be considerably less than
0. 1%. ff pinoresinol structures were present in lignin in appreciable amounts, one
had to assume that they are linked to adjacent units by acid-stable bonds, i. e.,
biphenyl and diaryl ether bonds, to an unexpectedly great extent. The 13C NMR
spectrum of spruce lignin [Nimz et al. 1974] also indicates a very low content of
pinoresinol structures. Experimental evidence in favor of the occurrence of such
structures in conifer lignin has been presented, however [Freudenberg et al. 1965;
Ogiyama, Kondo 1966, 1967].
A further acidolysis product exhibiting/~-/3 coupling, namely D,L-3,4-divanillyltetrahydrofuran (18. 7), was isolated in small amounts from spruce lignin acidolysis
[Lundquist 1970]. It can be assumed to originate from structures of type 23.4 present in lignin. The compound differs from the other dimeric acidolysis products in
posses~rtg a lower degree of oxidation. Its formation seems to involve an oxidoreduction process. Degradation of beech wood by thioacetolysis [Nimz 1974]
(p. 197) afforded the syringyl analog of compound 18. 7.
Finally, the 1,2-diarylpropane-l,3-diol structure (23.5) and the glyceraldehyde2-aryl ether structure (23.6) containing the side chain detached from one of the
units incorporated into the former structure have been derived from acidolysis
studies.

MiM hydro~s~
With the aim of finding a method which would degrade lignin without involving
simultaneous condensation reactions, Nimz [1966b, 1974] subjected preextracted
spruce and beech wood to percolation with water or 2 % aqueous acetic acid at
100 ~ for Several weeks ('~mild hydrolysis"). The extracts afforded D,L-guaiacylglycerol (16.1; 0.04% of the lignin of spruce wood) [Nimz I967] and D,L-syringytglycerol (2.5 % of the lignin of beech wood) and a group of eight dimers in addition
to two diastereomeric trimers and one tetramerie compound.
The dimers included D,L-syringaresinol (18.8; 2 %, beech) [Nimz, Gaber 1965],
as well as guaiacylglycerol-~3-coniferyl ether (24.1; 0.05 %) and similar/3-aryl ethers,
in which the coniferyl alcohol moiety was replaced by a confferaldehyde group, a
guaiacylglycerol group (0.04 %), or a vanfllin group. Finally, three dimers indicated
by formula 24.2 were isolated, one of them containing two guaiacyl residues
(0.12 %, spruce), the second with a guaiacyl residue in the a-position and a syringyl
residue in the ~-position of the side-chain (0.1%, beech), and the third one with
two syringyl residues (1.5 %, beech).
The modes of interco~mections between C 9 units present in these degradation
products have been revealed indepedently by the acidolysis studies discussed in the
preceding section. New features introduced by the "mild hydrolysis" of Nimz were

194

E. Adler

the free glycerol sidechains present in guaiacylglycerol and syringylglycerol, as well


as in the guaiacylglycerol-fl-guaiacylgtycerol ether (24.1, with CHOH-CHOH-CH20H
instead of CH=CH-CH2OH) mentioned above.

Mitd Hydrolysis of Wood by Percolation with Water


cat 100"
CH20H

CH
~H

H2~OH [ ~ O C H 3

H2((OH

..'OcH3

! CH30..'~CH3

OH

OH

Fig. 24.

It has been assumed that the monomeric arylglycerols are liberated by hydrolytic
cleavage of the ~-aryl ether bond in structures of type 24.1 [Wallis 1971]. Higuchi
et al. [1974] showed, however, that guaiacylglycerol is not produced when the model
compound guaiacylglycerol~-guaiacyl ether (16.3) is treated under the conditions of
Nimz' "mild hydrolysis". It was further reported by the same authors that guaiacylglycerol, as well as its syringyl and p-hydroxyphenyl analogs, are formed in yields of
0.03-0.6 % on enzymatic dehydrogenation of coniferyl alcohol and its corresponding
analogs (Fig. 8). Therefore, it was concluded that units carrying free glycerol sidechains were native end groups in li~in.
The liberation of the phenolic compounds obtained is assumed by Nimz to be
due to cleavage of easily hydrolyzable benzyl aryl ether linkages (formula 24.3), the
~-aryl ether linkages largely remaining intact. Secondary reactions involving/3-aryl
ether cleavage cannot be entirely excluded, however. In a model experiment, Nimz
[1966 c] found that the dihydro derivative of guaiacylglycerol-~-coniferyl ether
(24.1, CH2CH2CHzOH instead of CH=CH-CH2OH), when heated in water for 7 days
at 100 ~ gave, in addition to unknown products, dihydroconfferyl alcohol (41%)
and the dihydro derivative of phenylcoumaran 9.2 (32%): A possible mechanism
for this unexpected reaction will be discussed in the following section.

Hydrolysis with dioxane-water at 180 ~


Sakakibara and co-workers heated various wood species and Bj6rkman hgnin with
water-dioxane (1:1) at 180~ for 20 minutes and demonstrated by chromatography
or, in some cases, by isolation the formation of low-molecular phenols, most of
them identical with degradation products found earlier by Nimz on "mild hydrolysis".

Ligninchemistry-past, present and future

195

They detected, for instance, guaiacylglycerol (16.1) and p-hydroxyphenylglycerol


[Sano, Sakakibara 1970a], syringaresinol (18.8) [Omori, Sakakibara 1971], 1,2-diguaiacyl-l,3-propanediol (20.1) and a trimer, in which 20.1 was linked by a/3-aryl
ether bond to a guaiacylglycerol unit [Sano, Snknkibara 1970].
Sak~kibara further found coniferyl alcohol and its analogs (Fig. 8) among the
hydrolysis products [Sakakibara, Nakayama 1962]. It was believed that these cinnamyl alcohols were end groups linked to the lignin by a hydrolyzable benzyl aryl
ether bond (cf. Fig. 24). When it was found that, remarkably enough, a model
compound, guaiacylglycerol-~-guaiacyl ether (16.3), under the conditions of the
high-temperature hydrolysis also afforded coniferyl alcohol (in addition to guaiacol
and other products), the corresponding structures in lignin had to be considered as
an additional source of the coniferyl alcohol [Sakakibara et al. 1966].
The hydrolysis of model compound 16.3 (180 ~ 20 minutes) was studied in
detail by Sano [1975]. Surprisingly, he detected, in addition to some starting material (24.0 %), not only coniferyl alcohol (2.6 %) but also pinoresinol (9.3, 0,8 %),
1,2-diguaiacyl-l,3-propanediol (20.1, 1.4 %), dehydrodiconiferyl alcohol (9.2, 4.5 %)
and two "trimeric" compounds, one of which consisted of a phenylcoumaran and
a t-ether moiety (type 19.1, 3,8 %), whereas the second contained two/3-aryl ether
links (3.6 %). Finally, the formation of substantial amounts of polymerized material
was observed.
Undoubtedly, the product pattern suggests that the reaction imitates the dehydrogenative formation of lignols and higher polymers from coniferyl alcohol. Sano
proposes that under the conditions used the model compound 16.3 looses a molecule of water to give the corresponding quinonemethide, followed by homolytic
cleavage of its aryl ether bond with the formation of a coniferyl alcohol radical
(cf. Fig. 11) and a guaiacoxy radical. Dimetizarion of the former radical will give
rise to the dilignols obtained, whereas a radical transfer reaction between this radical
and the phenol 16.3 will give coniferyl alcohol and a new phenoxy radical. "Endwise" addition of a coniferyl alcohol radical to the latter phenoxy radical accounts
for the formation of the trimers, and, similarly, radical transfer and coupling reactions can result in higher-molecular products.
This unexpected multitude of reaction products arising from the fl-aryl ether
model 16.3 raises the question whether coniferyl alcohol and aroxyl radicals generated by homolytic cleavage of/~-aryl ether bonds contribute to the formation of
some of the low-molecular compounds obtained in the high-temperature hydrolysis
of lignin. The model experiment [Nimz 1966 c] mentioned at the concluding part
of the preceding section (p. 194) seems to indicate that similar reactions may also
occur in Nimz' "mild hydrolysis". It is very probable, however, that in both procedures the hydrolytic cleavage of benzyl aryl ether linkages is the major reaction.
It may be pointed out that, in contrast to the hydrolytic procedures according to
Nimz and Sakakibara, acidolysis of the fl-aryl ether model 16.3 gives a clear-cut
picture of the degradation of arylglycerol~-aryl ether structures (cf. p. 187), no
complicating homolytic cleavage with subsequent radical recombinations tzking

196

E. Adler

place under these conditions. Obviously, initial formation of the benzylium ion
brought about by the high concentration of protons is responsible for the formation
of an enol aryl ether intermediate, the latter undergoing heterolytic rather than
homolytic cleavage [Adler, Pepper, Ericsoo 1957].

Thioacetoly~s
In addition to the method of "mild hydrolysis", Nimz [1969, 1974] has introduced
a degradation procedure, called thioacetolysis, which causes cleavage of/3-0-4 bonds
and therefore brings about a deep-going fragmentation of the lignin. As much as
91% of the lignin of beech wood and 77 % of the lignin of spruce wood were degraded to mixtures of monometic to tetrameric products. The principle of the
three-step degradation method has been formulated by Nimz as shown in Fig. 25.

~H20H
HCOAr
I
HCOR"
R'

~H2OH
HCOAr
HL ~

OCH3
OR

R'

OCH3

R' ~ "OCH3
OR
3

OR
2

INzOCOCH3
HCOAr
I
~OCH 3

H\

~H20H
HCSI
/

HC-

/.~OC
R'o ~ O C H /
OR

/.

Fig. 25.

CH20H
cH2
CH2

OR
6

Ni[HI
n

R'/~OCH3
OR

Degradation of lignin with thioacetic acid. (From Nimz 11969])

Treatment of wood with thioacefic acid and boron trifluofide converts the arylglycerol~-aryl ether trait 25.1 via the benzylium ion 25.2 into the S-benzyl thioacetate 25.3. Subsequent saponification with 2N NaOH at 60~ gives a benzyl
thiolate ion (25.4) which looses the/3-aryloxy group by nucleophilic attack of the
neighboring thiolate ion on the/3-carbon atom to give an episulfide (25.5). The
latter dimerizes to dithianes or potymerizes to thioethers of type 25.6 [cf. Gierer
et al. 1964; Gierer 1970]. In a final step, treatment with Raney nickel and alkali
at 115 ~ removes the sulfur and yields the reduced phenolic reaction products
(R in formulae 5 - 7 denotes H).
In the monomer fractions from spruce and beech, 4-propylguaiacol and 4-propyl2,6-dimethoxyphenol, respectively, predominated, while dihydroconiferyl alcohol

197

Lignin chemistry-past, present and future

(25. 7, R = R' = H) and dihydrosinapyl

alcohol (25. 7, R = H, R' = CH30), respectively,


were formed in smaller amounts [ N m z et al. 1971]. In addition to the conversion
of 7-hydroxymethyl groups into methyl groups, loss of the former groups, probably
as formaldehyde, also takes place, "dimeric" products carrying side-chains with two
carbon atoms being frequent. The twenty "dimers" obtained from beech wood are
H2~R'

CH~I

CH~

CH2

X
~-

-oc~

"F

OH
R

-oc~

~--f-

OH

-oc~

R'-H

OH

RsR'=H

R = OCH 3,R'= H

R = OCH3,R'= H

R - IR'= OCH 3

R ~ H~R'~ OH
y1,2 ~

3~

~ H3

CH3

HC

CH31

~ H3

He/-~-

CR'

OH

OH

CH

o~

OH

CH 2

CR'

OH

OH

H~ /

OH

CH 2

OH

R=H,R'=H 2

OCH 3 ,R'= H 2

R = OCHs~R'= 0

R = OCH 3 ~R'- H 2

1%

R = OCH 3 ~R'- O
Q'5O/o

CH3

CH 2

CH 2

CHv

CH2

OH

OH

OH

R OCHs~R'=H
R = OCH 3 ,R'= OH

OH

R= OCH3

o,z.5 ~

o/, ~

CH 2

OH
R=OCHs~R' H
R = OCHs,R'- OH

%3~

HC~

CH

OH

OH

R-OCH 3

oJ %

Fig. 26. Dilignols obtained on degradation of beech lignin with thioacetic acid. The yields of
analyti~lly pure products (mean values of several determinations) given under the formulas are
based on lignin. (From Nimz [1974])

198

E. Adler

shown in Fig. 26 [Nimz, Das 1971; Nimz 1974]. Most of the bond types exhibited
by these dimers are identical with those revealed by other degradation methods,
especially acidolysis (Fig. 23) and oxidative degradation (p. 200).
The a-/3-junction found in some compounds (0,5 %) was assumed by Nimz and
Das [1971] to be present in beech lignin, although it cannot be the result of a
dehydrogenation. Its formation and that of some related structures is assumed
[Glasser, Glasser, Nimz 1976] to be due to a proton-catalyzed polymerization of
coniferyl alcohol or coniferyl alcohol end groups caused by the natural acidity of
the cell sap [cf. Freudenberg, Maercker, Nimz 1964]. Miksche and Yasuda [1976]
however point to the possibility also mentioned by Nimz [1969] that coniferyl
alcohol and sinapyl alcohol may arise during the degradation process, in analogy to
the formation of coniferyl alcohol as an intermediate in kraft cooking of spruce
lignin [Brunow, Miksche 1972; Gierer et al. 1973]. The heating with alkali in the
last degradation step may cause Michael type additions of the/3-C atom of coniferyl
alcohol to the a-position of quinonemethide structures which also can be assumed
to be intermediates. An analogous ~-3' addition has been found in the alkaline
dimerization of coniferyl alcohol [Aminoff et al. 1974; Brunow, Miksche 1976].
The a-~ linked products may accordingly be artefacts [Miksche, Yasuda 1976].
On the basis of the yields of crude and pure degradation products, Nimz [1974]
has calculated the frequencies of the various bond types in beech lignin and has
also proposed a structural scheme for this lignin (p. 207).

Oxidative degradation
The earlier, fundamental work on the permanganate oxidation of methylated lignin
[Freudenberg et al. 1936, 1938] (13. 173) was continued in studies using distribution
between solvents and column chromatography which resulted in the isolation of
nineteen methoxyl-substituted benzene carboxylic acids [Freudenberg et al. 1962,
1967; Freudenberg. In: Freudenberg, Neish 1968, p. 78]. In this work, the permanganate oxidation was carried out at pH 6-7. Miksche and co-workers [Larsson,
Miksche 1967, 1969 a] found that considerably higher yields of the aromatic carboxylic acids were obtained if the oxidation was carried out in aqueous sodium carbonate. Since the product mixture contained appreciable amounts of phenylglyoxylic
acids (27.4), however, the latter acids were degraded to the corresponding benzoic
acids in a second oxidation step, consisting of brief treatment with alkaline hydrogen peroxide. The mixture of benzoic acids (27.5) was finally methylated, and the
resulting mixture of methyl esters (27.6) subjected to gas chromatography. This
permitted the structural identification by mass spectrometry of individual fractions
appearing in the chromatogram, as well as the quantitative estimation of the more
prominent constituents [Larsson, Miksche 1969a]. For the identification of trace
components, the mixture was fractionated by column chromatography prior to the
gas chromatographic examination [Larsson, Miksche 1969a, 1971b, 1972].

Lignin chemistry-past, present and future

199

Oxidative Degradcltion of Lignin

C-C-C

C-C-C

(0.3o)2so2

"

OMe

pill1
OH
OH-(Kroft cook)

(1)
[ C0[~ FI
KMnO~/NaIOZll
+
0Me' OH,-82'
...
OMe
t-~OH
oM~OMe
3

~00H
CO
OMe

(2) 1 ~ 0 2 , pH 9-10,50 ~

C-C-C

C-O

Fig. 27.

00Me
6 "~'O~.. 4CH2N2
'" ~r- 0Me
0Me
gos chromotographic
separotion ond
estirnotion

COON

OMo
5

If wood or isolated lignin were methylated and oxidized, the resulting aromatic
acids reflected the units in lignin which carried a free phenolic hydroxyl group
(27.2). Preheating with alkali converted ethefified units in the lignin (27.1) into
phenolic ones, and from the increase in the yield of aromatic acids, therefore, the
proportion of etherified units could be estimated [Larsson, Miksche 1971a]. Alkaline ether cleavage was generally performed under kraft cooking conditions, but
higher yields of aromatic acids were obtained by treatment with hot aqueous sodium
hydroxide in the presence of cupric oxide according to the procedure used by Pearl
[1942].
The fin-st step was later improved in that the permanganate was replaced by a
mixture of sodium periodate and permanganate [cf. Lemieux, yon Rudloff 1955].
The oxidation was carried out at 82 ~ in aqueous sodium hydroxide in the presence of t-butanol [Erickson et al. 1973a]. The method gave well reproducible
results.
A total of forty aromatic acids has been identified in the reaction mixture obtained on oxidation of methylated spruce Bj6rkman lignin which had not been subjected to ether cleavage [Larsson, Miksche 1969a, b, 1971b, 1972]. A selection of
17 degradation acids is given in Fig. 28, which shows the most abundant acids obtained from softwoods (28.1-28.5, 28. 7 and 28.9) and from hardwoods (28.228.9). The remaining adds, such as 28.10-28.17, appear in small amounts (around
0.1% of the lignin) or in traces (< 0.1%). Compounds 28.11-28.17 were not
among the nineteen aromatic acids reported by Freudenberg et al.
The diaryl ether 28.10 and the biphenyl compound 28.11, being monocarboxylic
acids, originate from substructures, from which one of the sidechains has been
detached (cf. p. 190). The tricarboxylic acid 28.12 is one of the examples indicating a mixed radical coupling (coniferyl and p-coumaryl alcohol) and, analogously,
the diaryl ether 28. 8 is derived from a substructure formed from a ~napyl and a
coniferyl radical. The trimethoxylated ring in the trace constituents 28.15-28.17

200

E. Adler

SomeOeqradotionAcids
COOH

COOH

~OCH 3

oc~

COOH

COOH

COOH

COOH

0CH3
3

OCHa
4

0CH3
5
~ H

0CH3
5

COOH

cOOH
d~
COOH
CH30

CCH3
O ~

OCH3
oc~

~H3
8

COOH

COOH
CH30

/--<OCH3 C H g O ~ O C H ?
X~/'--OCH3
OCH3

CHsO~~
~ "(::~+3
OCH3 0CH3

Io

COOH

COOH

OCH3

COOH
OCH3

CH30

11

12

COOH

COOH

COOH
COOH

CH30"~~/OCH30CH
3 0
~..OCH3

CH30/~

A 00~

0-~0/0CH3

0CH3

C00H
~OCH 3
OCH3

/4

COOH 0CH3
CH30

OCH3 CH30
OCH3
O-

OCH3

~s

FiB. 28.

CH30~OCH3
OCH3 0CH3

OCH3

is

is remarkable. It points to the presence in lignin of a methoxyhydroquinone structure. The latter may have been formed by reduction of a methoxy-p-quinone
moiety which, in turn, may be the hydrolysis product of a 2,4-cyclohexadienone
diaryl ketal structure (Fig. 29, K). The latter type of structure may well arise by
dehydrogenative coupling [Larsson, Miksche 1972].
Table I shows the yields of the major acids obtained on degradation of a methylated spruce Bj6rkman lignin and of the same lignin which had been subjected to
ether cleavage prior to the methylation. An experiment with methylated kraft lignin
prepared from spruce wood meal is also included [Erickson et al. 1973b].
The parallel experiments 1 and 2, and 4--6, respectively, show the good reproducibility of the method. Comparison of the latter experiments with experiment 7

Lignin chemistry-past, present and future

201

Table 1, Yields of methyl esters (in rng per 100 mg of lignin) obtained on oxidative degradation
of methylated spruce lignin (for details, see Erickson et al. I1973b])
Expt.
Nr.

Methyl ester of acid (rag/100 nag of lignin)


28.1
28.2
28.3
28.4

28.5

28. 7

28.9

0.75
0.85

1.1
1.1

1.55
1.55

2.1

6.0

Bj6rkman lignin, methylated


1
2

0.7
0.65

11.2
10.7

0.25
0.25

2.0
2.0

Bj6rkman lignin heated with 2 M NaOH/CuO a and methylated


3

0.7

29.8

0.5

5.0

1.t

Bj6rkman lignin treated under kraft cooking conditionsb and methylated


4
5
6

0.5
0.5
0.5

21.4
20.6
21.1

0.15
0.15
0.15

6.1
6.0
5.8

0.55
0.55
0.55

1.75
1.75
1.6

3.6
3.9
4.0

0.6

1.65

3.4

Kraft lignin from wood meal, methylatedc


7

0.5

21.3

0.45

5.9

a 2hours, 170~
b 3.5 g of NaOH and 3.1 g of Na2S, 9 H20 in 100 ml of water, 3 hours, 170 ~
c Yields in rag per 400 mg of wood
indicates that in the case o f spruce the Bj6rkman li~,nin is structurally very similar
to the lignln in the wood.
Pretreatment with NaOH/CuO gave considerably higher yields o f most o f the
acids (expt. 3) than the pretreatment under kraft cooking conditions (expts. 4~5).
This is assumed to be due to a favorable "preoxidation" o f the side-chains by the
NaOH/CuO system, taking place in addition to ether cleavage.

The

frequencies o f

the different bond types

Spruce lignin

The yields o f oxidation products given in experiments 1 and 2 (mean value) and in
experiment 3 (Table I) were used in an estimation o f the frequencies o f the different types o f C 9 units corresponding to the degradation acids. The former values
refer to the phenolic portion, the latter to the total o f originally phenolic and
ethetified units in spruce Bj6rkman lignin.
On the basis o f the frequencies o f the different types o f phenylpropane units
and the number o f phenolic hydroxyl groups liberated on treatment with NaOH/
CuO, estimates could now be made of the proportions o f the types o f bonds interconnecting the C 9 units in spruce lignin. They are collected in Table 2; in a few
cases, bond type frequencies were derived from other results thzrt those obtained

202

E. Adler

on oxidative degradation. The bond types mentioned in the Table are shown in
Fig. 29.
In the calculations leading to the estimates given in Table 2, certain assumptions,
e.g. regarding the actual and theoretical yields of the oxidation products, are involved. For these details, the reader is referred to the original papers [Larsson,
Miksche 1971a; Erickson et al. 1973b].
Table 2. Proportions of different types of bonds connecting phenylpropane units in Bj6rkman
lignin from spruce (Plcea abtes). Letters A - K refer to Fig. 29
Bond type

Proportions
%

Basis of the estimate

A (in arylglycerol-r
ether
structures)
B (in glyceraldehyde-2-aryl ether
structures)
C (in noncyclic benzyl aryl ether
structures)
D (in phenylcoumaran structures)

48
2

Liberated phenolic OH minus structures


B,C,D
Acidolysis; 1HNMR (p. 191)

6- 8

cf. p. 184 [Adler et al. 1968]

9-12

Upper limit from 28.4 in expt. 3


(Table 1); lower limit, c2".Adler, Lundquist [1963]
28.5 in expt. 3 (Table 1)a

E (in structures condensed in 2- or


6-positions)
F (in biphenyl structures)

2.5- 3

G (in diphenyl ether structures)

3.5- 4

H (in 1,2-diarylpropane structures)


J (in fl-~ linked structures)
K (in quinone ketal structures) d

9.5-11

7
2
traces

Lower limit from 28.9 in expt. 3


(Table 1); upper limit includes trace
constituents
Lower limit from 28. 7 in expt. 3
(Table 1); upper limit includes trace
constituents
Acidolysisb
AcidolysisC; cf. also pp. 192-193
cf. Larsson, Miksche [1972]

a 2-Condensed structures occur only in very small amounts, hemipinic acid being a trace constituent [Larsson, Miksche 1969b].
b Estimate based on yields of acidolysis products [Lundquist 1970].
c In acidolysis, pinoresinol was not found [Lundquist 1970]; 13C NMR data also indicate a
very low content of this structure [Nimz et at. 1974]. However, D,L-divanillyltetrahydrofuran (18.7) was obtained on acidolysis of spruce lignin (c.f.p. 193).
d Alternatively, the corresponding p-benzoquinone or hydroquinone structures (cf. p. 200).
Naturally, the frequency values given in Table 2 are to be regarded as approximate rather than accurate, although most of them probably are fairly correct. It
should also be noted that the experiments were carried out with Bj6rkman lignin.
This implies that the value given for bond type C may be somewhat below the unknown amount of C in the original lignin of the wood (of. p. 183). The figure
given for the most prominent bond type (A) may then include a minor portion of
bonds of type C which have been opened during the milling procedure.

203

Lignin chemistry-past, present and future


C
~-0~--c

C
B

F
C
I
C

0
0
F (traces)

F (troces)

6
Ar0~0A r
0

0
G
FiB. 29.

0
G~races~

Prominent Structures in Softwood Lio#nin

6
~

HCI=O[CI-120H]
HC
CH30

~H

~H20H
HC-O-I

?.2o.
HC-OCH

HCOH

cH~

cH

H
O--~CH
I
HCOH

HOCH2

&
&o~

,, -?, ~ .

oc~

CH2OH HCI
HO-~CH
O'CH3HCOH

CH
I

i~o~

~o.~cc~o~;-

HOCH2 I ~

~C@ocm

HCI
H -OH

HC---O
CH30~
O

F ~ 30.

ICH2OH
CH
I
HCOH

OH

mcoHc ~
He
I

OCH3
OH[~-C]

CH3

204

E. Adler

Some uncertainty is attached to the figure given for bond type H. It was assumed earlier [Erickson et al. 1973] that the frequency of structure H equals that
of B with which it is genetically connected and which has been determined analytically (p. 190). This would give a frequency of 2 % for bond type H. However, an
evaluation of the yields of acidolysis products originating from structures of this
type [Lundquist 1970] gives an estimated value of about 7 % for H.

A structural scheme for spruce lignin


The prominent substructures of spruce lignin are collected in a scheme (Fig. 30)
comprising 16 C9 units. The sequence of the different units has been chosen arbitrarily. Furthermore, in a scheme of that limited size the proportions of certain
structural details cannot be accounted for in a correct way. For example, the
amount of syringyl units in spruce ]ignin has been estimated to be 1% (cf. p. 181);
thus, the incorporation of one syringyl ring (nr. 13) into the scheme has a qualitative
rather than a quantitative meaning. Furthermore, the presence of one pinoresinol
entity (hrs. 10 and 11) probably overemphasizes the proportion of this substructure
(cfi p. 193). In addition, at least part of the fl-fl linked substructures is of the divanillyltetrahydrofuran type (18. 7) rather then the pinoresinol type.
A number of minor or trace structures is recorded in Fig. 31. Structures 1-11
are derived from trace constituents detected on oxidative degradation of methylated
Minor 5tructurol Constituents o! Softwood Li~nln

i
. ~ ~

~ / T ~ CH3
CH30~

0
0

C
and~or

OCH3

"",,,~OCH3
2 0

~OCH
0 4a

-C-C

-~-C
OCH3

CH30

O~3
OH

OCH3

C-C-C OH

cH3~ocN
0
OH

&

O5
C

CH3

O 5Q
H~H2OH CH2OH
"CH

~H2 'H2

~ O C H 3 CH30~ ~OCH30
13
/2
F~. 31.

Og
H2C'~)~ (~H2
H~H

C
7

OHpo

CH
2 H~

GH3c~O
7~;

OCH3 CH30

CH3

CH3

205

Lignin c h e m i s t r y - p a s t , present and future

spruce lignin. The cyclolignan structure 11 would explain the oxidative formation
of benzenepentacarboxylic acid [Read, Purves 1952; Brunow 1967]. Traces of
eugenolglycol have been detected among the low-molecular products formed in soda
cooking of spruce Bj6rkman lignin. Although it cannot be excluded that the product is formed from a known structure, the presence of traces of structure 12 in
lignin may be considered [Lundquist 1973]. The reduced benzyl position would be
parallelled by the occurrence of structure 14, demonstrated by acidolysis. Since
compound 13 (OH instead of O) under acidolysis conditions is converted into 14
(OH instead of O), it cannot be decided whether 13 or 14 is the precursor of the
latter compound.

Lignin in compression wood


It is well-known that compression wood is characterized by an abnormally high
degree of lignification [C6t6 et al. 1966]. The unusually low methoxyl content of
the lignin is due to an increased incorporation of phenylpropane units of the
p-hydroxyphenyl type, as indicated by the formation of comparatively high proportions of Hibbert ketones of this type [Latif 1968] and of p-hydroxybenzaldehyde
on nitrobenzene oxidation [Bland 1961].
Degradation of compression wood and normal wood from P/nus mugo by the
two-step oxidation procedure described above revealed that the former wood gave
strongly increased yields of anisic acid (28. I) and 4-methoxyisophthalic acid as well
as of diaryl ether and biphenyl carboxylic acids containing one or two rings derived
from p-hydroxyphenyl units. Several trace constituents of these types were detected
in the case of the compression wood only [Erickson et at. 1973c].
The results indicated that the compression wood lignin contained about 20 phydroxyphenylpropane units per 100 C9 units, i. e., about four times more than the
normal wood. Some of the substructures characteristic of compression wood lignin
are shown in Fig. 32.
Substructures

Typical for Compression Wood


( Pmu5 mug~)

C-C-C

C-C-C

C-C-C

C-C-C

0
l

C-C-C

C-C--C

0
2

C-C--C

c-c-c

C-C-C

c_c_ o

C-C-C

FiB. 32.

,~

206

E. Adler

The "cross-coupled" structures 3 and 4 are o f special interest because they


demonstrate the copolymerization o f guaiacyl and p-hydroxyphenyl components.
Structures 5 and 6 belong to the trace structures which have been found only in
compression wood.

Hardwood lignins
Table 3 shows approximate values for the frequency of different bond types in
birch lignin. Most o f them were derived from the yields o f acids 2 8 . 2 - 2 8 . 9 obtained on oxidative degradation of the methylated lignin, without and with pretreatment with NaOH/CuO [Larsson, Miksche 1971 a]. The figures for b o n d types B, C,
H and J are based upon results from acidolysis studies.
Table 3. Frequencies of different types of bonds connecting phenylpropane units in Bj6rkman
lignin from birch (Betula verrucoga). Letters A - J refer to Fig. 29
Bond type
A (arylglycerol-~-aryl ether)
B (glyceraldehyde-2-aryl ether)
C (noncyclic benzyl aryl ether)
D (phenylcoumamn)
E (in 2- or 6-position condensed
structures)
F (biphenyl)
G (diphenyl ether)
H (1,2-diarylpropane-1,3-diol)
J ~ linked structures)

Frequency (per 100 C 9 units)


Guaiacyl type Syringyl type
22-28

1-1.5
4.5
1

34-39

0.5-1
5.5

Total

Remarks

60
2
6-8
6

cf. p. 190
a

1.5-2.5
4.5
6.5
7
3

b
c

a The same value as that determined for spruce lignin ]Adler et al. 1968] was assumed. "Mild
acidolysis", which hydrolyzes noncyclic benzyl azyl ethers (cf. p. 184), brought about a degradation of the birch lignin similar to that observed with spruce lignin (gel filtration, [Larsson,
Miksche 1971a].
b Estimate based on yields of acidoiysis products derived from structure H [Lundquist 1973].
c Estimate based on yields of syringaresinol and episyringaresinol obtained on acidolysis
(cf. p. 192).
The yields o f degradation acids o f the guaiacyl and the syringyl type indicate a
guaiacyl : syringyl ratio of about 1 : 1 for the Bj6rkman lignin o f birch. As seen
from Table 3, the syringylpropane units are etherified as/Laryl ethers to a considerably higher degree than the guaiacylpropane units and in the phenolic portion of
the lignin the guaiacyl : syringyl ratio is accordingly 1 : 0.36.
F r o m the relative yields o f monomeric and dimeric degradation products isolated
after thioacetolysis o f beech wood, Nimz [ 1974] calculated the proportions of
various bond types in beech lignin (Table 4).

Lignin chemistry-past, present and future

207

Table 4. Proportions of various bond types in beech lignin [Nimz 1974]. Letters A-J refer to
Fig. 29
Bond type

Proportions
%

Acyclic a- and/~-ethers (C + A)
#-1 bonds (H)
~-5 bonds (D)
bonds (syringaresinol and
pinoresinol units, J)
a-~ bonds
5-5 bonds (F)
bonds in d~enzyRetrahydrofuran units (J)
5-0-4 bonds (G)
and a-6 bonds (tetralin units)

65
15
6
5
2.5
2.3
2
1.5
0.5

The general picture obtained by Nimz is similar to that presented in Table 3 for
birch lignin. Apart from the fact that the two sets of figures refer to lignins from
different hardwoods, Bj6rkman lignin was used in the case of birch, and the total
lignin of the wood in the case of beech. These differences and, in addition, the
different experimental procedures used, may explain some of the dissimlqarities
between the two sets of data [cf. also Miksche, Yasuda 1976].
From the point of view of the recent results of Goring et al. concerning the uneven distribution of guaiacylpropane and syringylpropane units over different cell
regions of hardwoods, any presentation of the composition of a hardwood lignin
refers to an average lignin, which pa" se does not exist in the wood. This must be
borne in mind when one is talking about the elemental composition and the
OCH3/C9 ratio or the frequencies of various substructures in hardwood lignins as
those given in Tables 3 and 4, and when a structural scheme for a hardwood hgnin
is presented [see Nimz 1974]. Some results of Goring's work are briefly summarized
in the following chapter.

The distrfbution of lignin in wood tissues and the location of guaiacyl and syringyt
units

Goring and his group [Scott et al. 1969] improved the technique of ultraviolet
microscopy earlier applied to wood by Lange [1954]. UV light tran~xtitted through
ultrathin sections (0.5 #m) of wood was measured within the wavelength range of
lignin absorption (240-320nm).
From the absorbances and the cell wall dimensions, the concentrations of lignin
in the earlywood of black spruce were found to be 22.5 %, 49.7 % and 84.8 %
(w/w) for the secondary wall, the compound middle lameUa and the cell comers,
respectively. The secondary wall occupies 87.4 % of the volume of the tracheid

208

E. Adler

Fig. 33. Cross section of Epon embedded tracheids of black spruce earlywood photographed
in ullxaviolet light of wavelength 240 nm. The densitometer tracing was taken across the tracheid wall on the dotted line. (From Fergus et aI. 119691)

tissue and accomodates 72.1% of the total lignin, despite of the fact that the concentration of the lignin is low in this part of the cell. Although the lignin concentration in the middle lamella and in the cell comers is high, these regions, which
occupy only 12.6 % of the tissue volume, accomodate only 27.9 % of the total lignin.
The lignin distribution in birch fibers offers a rather similar picture [Fergus, Goring
1970a].
Whereas simple guaiacylpropane model compounds have an absorption maximum
at 280-282 nm with an extinction coefficient e = 19 cm -1 1 g-1 (4-propylguaiacol),
corresponding syringyl compounds show Xmax at 270-273 nm with e about 6crn -1
1 g-1 (2,6-dimethoxy-4-propylphenol) [cf. Fergus, Goring 1970b; Aulin-Erdtman,
Sand~n 1968]. A pure syringyl lignin therefore would be expected to have an absorption maximum which is located at a lower wavelength than that of a pure
guaiacyl lignin and about three times weaker than the latter one. As already found
by Lange [1954], the maximum of the absorption curve of spruce fibers, as expected,
is at about 280nm. This has been confirmed by Fergus and Goring [1970b] for
the middle lamella as well as the secondary wall of spruce tracheids.
Musha and Goring [ 1975] furthermore found that, for a series of different hardwoods, the absorbance of the secondary wall in fibers decreases with increasing
CH30/C 9 ratio of the total lignin of the wood, and the position of the absorption
maximum shifts towards shorter wavelengths. In the cell corner region, however,
the decrease in absorption with increasing methoxyl content is much less pronounced,
and this is also true for the blue .~ift of the absorption maximum.

Lignin chemistry-past, present and future

209

The conclusions drawn from these measurements were that in hardwoods the
lignin in the secondary wall of the fibers has a high syringyl content and that the
syringyl : guaiacyl ratio in this region increases with increasing CH30/C 9 ratio of the
total lignin, and that the cell comer and middle lamella llgnins are of the guaiacyl
type, their syringyl content rising only slowly with increasing CH30/C 9 ratio. The
lJgnin in vessel walls is mainly a guaiacyl lignin, as far as hardwoods with a low or
medium methoxyl content are concerned; if the CH30/C 9 ratio becomes > 1.50,
the syringyl content of the vessel llgnin increases rapidly. Musha and Goring point
to the difficulty in determining the location of the absorption maximum, the absorption curves of lignins with high syringyl content being rather fiat. If it were
possible to separate the different cell regions and cell types from each other, the
correctness of the conclusions mentioned above could be tested.
It has been assumed that Bj6rkman llgrdn originates nminly from the middle
lamella. This cannot be the case, if Musha and Goring's view that middle lamella
lignin is mostly of the guaiacyl type is correct. Although Bj6rkman lignins from
hardwoods have a slightly lower syringyl : guaiacyl ratio than the total llgnin of the
corresponding wood [Larsson, Miksche 1971 a; Chang et al. 1975; Miksche, Yasuda
1976], the differences are small and the Bj6rkman llgnin preparations should contain a large proportion of secondary wall lignm.
Kirk et al. [1975] found that, during the initial phases of the decay of whole
birch wood by a white-rot fungus, syringyl units are removed preferentially, whereas
from ground wood both syringyl and guaiacyl units disappear at comparable rates.
The authors concluded that, in the case of whole wood, the fungus attacks from
the lumen and encounters first a sytingyl-rich llgnin. This would be expected from
the views advanced by Goring and co-workers.

Occurrence and composition of lis

in different plants

The distribution of lignius in the plant kingdom and their classification on a taxonomic basis has been excellently reviewed by Sarkanen and Hergert [1971]. These
authors favor the division of lignins into two classes only, namely "guaJacyl lignins"
and "guaiacyl-syringyl lignins" [Gibbs 1958]. Both types of lignin give a positive
Wiesner reaction, caused by cinnamaldehyde end groups, but only the latter exhibits
a positive M~iule reaction which is given by syringyi nuclei.
The Lignins of the majority of gyrrmosperms are typical guaiacyl lignins, although
they contain a small amount of syringyl units, not exceeding 1.5 % [Erickson,
Miksche 1974a], in addition to a minor proportion (about 5 %) of p-hydroxyphenyl
units. There are, however, exceptions comprising Ephedra and Gnetum species, as
well as a Tetraclinis and a Stangeria species, the hgnins of which have a high syringyl
content, similar to that found in angiosperms. Erickson and Miksche [1974] clearly
demonstrated that the lignins in these cases are true guaiacyl-syringyl ligrtins. Oxidative degradation (p. 198) gave the same degradation acids (28.2-28.9) as formed

210

E. Adler

from angiosperm lignins, and their yields were comparable to those obtained from
such lianins. Of special value was the detection of the diaryl ether 28.8 indicating
guaiacyl-syringyl copolymerization in the exceptional gymnosperm lignins. Podocarpus species show a moderately increased syringyl content (2-6 %) and, from this
point of view, are between the two types of lJ~ins.
Oxidative degradation was also applied in the characterization of angiosperm
lignins, as well as the hgnins of cryptogams. In the standard procedure, the lignin
was extracted from the plant material by kraft cooking (aqueous NaOH/Na2S, 170 ~
3 hours) and, after methylation and oxidation, the benzoic acids 28.1-28.9 were
determined as methyl esters by gas chromatography (p. 198). The lifimins of all
dicotyledon angiosperms investigated were found to be true guaiacyl-syringyl copolymers. According to IR spectroscopic investigations [Sarkanen et al. 1967] the
syringyl content of woody angiosperms varies between 20 and 60 %. In herbaceous
angiosperms (dicotyledons) this range was widened to 10~55 %, the average syringyl
content being lower than in the woody angiosperms [Erickson et al. 1973c].
Oxidation of the lignins of grasses (monocotyledon angiosperms) gives comparatively high yields of anisic acid (28.1) which might suggest the dehydrogenative incorporation of correspondingly large proportions of p-hydroxyphenyl units. It has
been shown, however, that coumaric and fenflic acids are bound as esters to the
grass hgnins [Higuchi et al. 1967]. Furthermore, the oxidation method revealed
that only small amounts of the binuclear acids originating from p-hydroxyphenyl
structures 3 Z 3 and 32.4, which are characteristic of compression wood (p. 206),
are formed from grasses [Erickson et al. 1973d; cf. also Yamasaki, Higuchi 1971].
The monocotyledon lJgnins investigated so far could therefore be classified as norreal guaiacyt-syringyl lJgnins.
Oxidative degradation of material extracted by kraft cooking from softwood and
hardwood bark [Andersson et al. 1973] largely confirmed the views which have
accumulated during recent years regarding the occurrence and chemical nature of
lignins in this part of the trees [Sarkanen, Hergert 1971]. The bark lignins from
conifers were guaiacyl li~rnirls which, however, seemed to contain a larger proportion
of p-hydroxyphenyl units than the lignins of the corresponding wood. Pre-extraction of the bark with 1% aqueous sodium hydroxide (90 ~ 1 h) which is expected
to remove polyphenolic material ("phenohc acid") resulted in a 20-30 % increase
in the yields of degradation acids. On oxidative degradation of the isolated "phenolic acid", the mononuclear carboxylic acids were formed in rather small yields,
but the binuclear products (28. 7 and 28. 9), which are characteristic of lignin, could
not be detected. Kraft lignins obtained from hardwood bark gave the degradation
acids expected for guaiacyl-syringyl h trains. Their syringyl content, however, was
considerably lower than that found for corresponding wood lignins.
With the exception of Selaginellales (four species examined), which contained
guaiacyl-syringyl hgnins, and a submerse aquatic plant (Isoetes lacustris) which was
iianin-free, all pteridophytes (vascular cryptogams) investigated contained normal
mmiacvt lilmin. The lj_anin content as i u ~ e d from the total yields of degradation

Lignin chemistry-past, present and future

211

acids, however, was low in a few cases, viz., in an Equisetum species and one of
seven species of ferns investigated.
Available evidence regarding the occurrence of lignins in bryophytes, which are
non-vascular cryptogams comprising mosses and liverworts, has been conflicting.
The very low methoxyl content of "Bj6rkman lignins" from the mosses Polytrichum
and Sphagnum [Freudenberg, Harkin 1964], the formation of p-hydroxybenzaldehyde on nitrobenzene oxidation of Sphagnum, as well as other data, suggested
the presence of a "p-hydroxyphenyl lignin" which was believed to be characteristic
of primitive plants [Freudenberg. In: Freudenberg, Neish 1968, p. 113-114]. However, ethanolysis of a moss species failed to give I-Iibbert ketones [Sarkanen, Latif,
quoted in: Sarkanen, Ludwig 1971, p. 81].
Oxidative degradation of six mosses and two liverworts [Efickson, Miksche
1974b] in each case yielded minor amounts of veratric, isohemipinic and metahemipinic acids. However, none of the binuclear acids (28. 7 and 28. 9), which would be
expected to be obtained from a lignin, could be detected among the oxidation products. Furthermore, Sphagnum gave anisic acid (28.1) in a considerable yield but
failed to give any other degradation acid of the p-hydroxyphenyl type. From these
results it was concluded that the Bryophyta species investigated did not contain any
hgnin.

On the linkage between li~in and hemicelluloses


In their discussion of the important problem concerning the existence of covalent
bonds between ljgnin and polysaccharides in the wood, Lai and Sarkanen [1971 ]
make the following statement:
"While the accumulated data rather convincingly support the existence of chemical bonds between polysaccharidic and lignin elements in woody tissues, the character of these bonds is not well understood." Although the problem is still far from
being solved, some advances, reported since Lai and Sarkanen's discussion was
published, may be mentioned.
Two major types of starting materials have been used in studies of lignin-carbohydrate complexes, namely holoceUulose and the water-soluble fraction obtained in
its preparation, or the lignin-carbohydrate complex (LCC) prepared from milled
wood according to Bj6rkman [1956, 1957].
From the water-soluble polysaccharides arising in chlorite treatment of spruce
wood, Kringstad and Cheng [1969] isolated a lignin-galactoglucomannan complex by
fractionation on Sephadex. After enzymatic hydrolysis, the remaining complex was
found to be enriched in galactose. Fengel [1976] and Fengel and Przyklenk [1975]
fractionated alkali extracts from spruce holocellulose by ion exchange chromatography and obtained two sets of hgnin-carbohydrate fractions, in which the ratio of
arabinose : lignin was 1 : 1 and 1 : 12, respectively.

212

E. Adler

Sephadex fractionation of LCC from pine wood followed by treatment of a


lignin-rich fraction with a/3-glucosidase gave small amounts of a product which was
soluble in ethanol-water (2 : 1) and contained 80 % lignin in addition to 9 % neutral
sugar units. Its molecular weight Mn was 3700 and it migrated electrophoretically
as a single spot. The results of methylation and hydrolysis were interpreted to indicate a glycosidic linkage of D-galactose, L-arabinose and D-xylose to the lignin
[Koshijima et al. 1976]. Eriksson and Lindgren [1977] similarly degraded spruce
LCC by means of a commercial ceUulase-hemicellulase and obtained a product which
after treatment with sodium borohydride contained 9 % monosaccharide and alditol
units. In this material, arabinose and galactose, which in softwood hemicelluloses
are present as monosaccharide side-chains [Timell 1967], were highly enriched.
All the results mentioned above seemed to suggest that galactose and arabinose
units act as links between lignin and hemiceUuloses. This suggestion received further
support from the identification of the alditol end groups in the borohydride-reduced
material of Eriksson and Lindgren. In a hydrolyzate from this material the authors
found xylitol, mannitol and glucitol and, interestingly, galactitol and arabitol. The
monosaccharides corresponding to the latter two alditols originally were bound, as
already mentioned, by glycosidic linkages to the glucomannan and xylan chains,
respectively. These linkages must have been hydrolyzed by the enzyme, but the
galactose and arabinose units remained in the enzyme-treated, lignin-rich material
and were precipitated together with the lignin after having been reduced to alditol
units. The finding that mild hydrolysis caused the formation of further amounts of
reducing arabinose and galactose units still bound to the lignin, led to the proposals
shown in Fig. 34 (upper left formula) and Fig. 35. Since mild hydrolysis removed
only about 60 % of the xylose and arabinose units, a second possible mode of linkage between lignln and arabinoxylan (Fig. 34, upper right formula) was also suggested.

Lignin

Lignin

Aro
--XyL- Xyl -- X.yt-

Mild

-- Xyt-Xyt--Xyl-

Ai-o

Aio A~a

Qcld

oci d
Ii Hitd
hydrolysis

hydrolysis

Lignin
i

Ara
--Xyt -XyL-- Xyl.-Aro

Lignin
]

--Xyt-Xyt--Xyl.-

Aro Aro

Fig. 34. Two suggestions for the linkages between lignin and xylan. Dotted lines are symbols
for furanosidic linkages. The uronic acid units in the xylan molecule are not written. (From
Eriksson and Lindgren [1977])

Lignin chemistry-pa~t, present and future

213

Lignin
]
Got
I

-Mon-Man-Gtu-Mon
i
i
I

Gat

MiLd acid

hydrolysis
Lignin
[

Got
-Man-Mon-Gtu-Men-,,
G~L
Fig. 35. A suggestion for the linkage between lignin and galactoglucomannan. Dashed lines
are symbols for the or-(1--,6)-pyranosidic linkages. (From Ert~sson and Lindgren [19771)
It has been shown in model experiments [Freudenberg, Grion 1959; Freudenberg,
Harkin 1960] that quinonemethides which are intermediates in the formation of
lignin (p. 179) are able to add carbohydrate hydroxyl groups. The phenolic benzyl
ethers thus formed are hydrolyzed by acids as easily as furanosidic linkages [Freudenberg. In: Freudenberg, Neish 1968, p. 93], for instance those by which arabinose
units are linked to xylan [Timell 1965]. Similar benzyl ethers carrying an etherified
phenolic hydroxyl group may arise by continued dehydrogenative polymerization
of the lignin and may be expected to be hydrolyzed less easily.

Outlook
The concept of lignin as a polymer formed by dehydrogenation of a mixture of
three p-hydroxycinnamyl alcohols is now well-established. The proportions of the
three monomers involved in the copolymerization process vary in different plants
and may vary even in different morphological parts of the plant, thus giving rise to
different lignins. In general, there are guaiacyl lignins, which are derived mainly
from coniferyl alcohol, and gualacyl-syringyl lignins, which are formed from coniferyl alcohol admixed with varying but substantial amounts of sinapyl alcohol. In
the formation of all lignins, a small amount (1-5 %) of p-coumaryl alcohol also
participates, and to our present knowledge, only the lignin of the compression wood
of conifers represents a true guaiacyl-p-hydroxyphenyl licrnin containing a substantial amount of units (about 20 %) derived from p-coumaryl alcohol.
The efforts to clarify the structures of the different types of lignin have resulted
in a detailed picture of the various modes in which the phenylpropane units are
linked together in the polymer. They are in harmony with the reactions expected
to take place on dehydrogenation of the p-hydroxycinnamyl alcohols. In addition

214

E. Adler

to the radical coupling reactions which follow the abstraction of a hydrogen atom
from the phenolic reactants, nondehydrogenative reactions, such as the addition of
phenolic groups to intermediary quinonemethides, are also involved.
Whereas there is good agreement regarding the frequency of the predominant type
of linkage, that is, the arylgtycerol-3-aryl ether structure, there is still some uncertainty regarding the proportions of some of the linkages occurring in minor amounts.
This is true, for instance, for the numbers of/3-1 coupled substructures and the
glyceraldehyde groups detached from them.
The balance of the different modes of substitution at the a-carbon atom, such as
the numbers of benzyl alcoholic and benzyl aryl ether groups, possibly also benzyl
alkyl ether groups, the keto function and the recently postulated a,/3-1LrLkages,also
needs to be further studied. Especially the numbers of benzyl aryl ethers and those
of free phenolic hydroxyl groups, as well as the amount of keto groups, are affected
by ball milling and, therefore, figures obtained with Bj6rknaan lignin will deviate to
a certain degree from those valid for the unchanged lignin in the wood.
Apart from these still necessary reffmements, our present knowledge regarding
lignin structure is probably fairly close to the truth, regarding both the qualitative
and the quantitative aspects. The attempts initiated by Glasser and Glasser [cf.
Glasser, Glasser, Nimz 1976 and earlier papers quoted therein] to analyse the correctness of available experimental data by computer simulation of the formation of softwood lignin seem to support this view.
The often discussed question of the existence of linkages between li~in and
carbohydrates in the wood now seems to be definitely answered in a positive sense
and some progress regarding their chemical nature has been made. A final solution
of these problems, however, will require much continued effort.
Goring's work on the heterogeneity of lignin in hardwoods naturally raises questions concerning the biological mechanisms by which the distribution of the coniferyl and sinapyl alcohols in the lignifying tissues is controlled.
In the past, interest in the behavior of lignin in pulping processes has greatly
furthered the development of its structural chemistry. To-day, the detailed insight
into lignin structure forms a solid basis for the clarification of the highly complex
reactions taking place in pulping and bleaching operations, and extensive knowledge
in these fields has already been gathered, mainly through the studies of Gierer
[1970], Gellerstedt [1976] and Miksche [Brunow, Miksche 1976].
Many attempts have been made to find application for the lignins from spent
pulping liquors. Although a number of them has been successful, the total amount
of lignin utilized commercially is only a small fraction of the millions of tons which
are extracted from wood during pulping and which mainly are used as fuel. It seems
reasonable to predict that efforts, directed towards economically feasible ways of
utilizing lignin, a renewable o~nnic material, will continue.

Lignin chemistry-past, present and future

215

References
Adler, E. 1957. Ind. Eng. Chem. 49:1377
Adler, E. 1959. Proc. IVth Internat. Congr. Biochem., Pergamon Press, London, Vol. 2:137
Adler, E. 1961a. Paperi Puu 43:634
Adler, E. 1961b. Das Papier 1961:604
Adler, E.; Becker, H.-D.; Ishihara, T.; Stamvl~, A. 1966. Holzforschung 2 0 : 3
Adler, E.; Bj6rkqvist, K. J. 1951. Acta Chem. Scand. 5:241
Adler, E.; Bj6rkqvist, K. J. 1953. Acta Chem. Seand. 7:561
Adler, E.; Bj6rkqvist, K. J.; Hiiggroth, S. 1948. Acta Chem. Scand. 2 : 9 3
Adler, E.; Delin, S.; Lundquist, K. 1959. Acta Chem. Scand. 13:2149
Adler, E.; EUmer, L. 1948. Acta Chem. Stand. 2 : 8 3 9
Adler, E.; Eriksoo, E. 1955. Acta Chem. Stand. 9:341
Adler, E.; Gierer, J. 1955. Acta Chem. Scand. 9 : 8 4
Adler, E.; Gierer, J. 1957. In: Tre~er, E. (Ed.). Die Chemic der PflanzenzeUwand. Springer,
Berlin, 457
Adler. E.; Gustafsson, B. 1963. Acta Chem. Scand. 17:27
Adler. E.; Hemestam, S. 1955. Acta Chem. Stand. 9 : 3 1 9
Adler. E.; Hernestam, S.; Walld~n, I. 1958. Sheens&Papperstid. 61:641
Adler E.; Lindgren, B. O. 1952. Svensk Papperstid. 55:563
Adler E.; Lindgren, B. O.; Saed6n, U. 1952. Svensk Papperstid. 55:245
Adler. E.; Lundquist, K. 1963. Acta Chem. Scand. 17:13
Adler. E.; Lundquist, K.; Miksche, G. E. 1966. Adv. Chem. Sex. 5 9 : 2 2
Adler E.; Marton, J. 1959. Acta Chem. Scand. 13:75
Adler. E.; Marton, J. 1961. Acta Chem. Stand. 15:357
Adler. E.; Miksche, G. E.; Johansson, B. 1968. Holzforschung 22:171
Adler. E.; Pepper, J. M.; Eriksoo, E. 1957. Ind. Eng. Chem. 49:1391
Adler. E.; YUner, S. 1952. Svensk Papperstidn. 55:238
Adler, E.; Yllner, S. 1953. Acta Chem. Scand. 7 : 5 7 0
Adler, E.; YUner, S. 1954. Svensk Papperstidn. 5 7 : 7 8
Aminoff, H.; Brunow, G.; Falck, K.; Miksche, G. E. 1974. Acta Chem. Scand. B 28:373
Andersson, A. A.; Erickson, M.; Fridh, H.; Miksche, G. E. 1973. Holzforschung 27:189
Aulin-Erdtman, G. 1954. Svensk. Papperstid. 57:171
Aulin-Erdtrnan, G.; Sand~n, R. 1968. Acta Chem. Scand. 22:1187
Becker, H.-D.; Adler, E. 1961. Acta Chem. Stand. 15:218
Berg, G. A.; Holmberg, B. 1935. Svensk Kern. Tidskr. 47:257
Berndtson, L.; Hedlund, K.; Hernr~t, L.; Lundquist, K. 1974. Acta Chem. Scand. B 28:333
Bj6rkman, A. 1956. Svensk Papperstidn. 59:477
Bj6rkman, A. 1957a. Svensk Papperstidn. 60:243
Bj6rkman, A. 1957b. Svensk Papperstidn. 60:329
Bj6rkman, A.; Person, B. 1957a. Svensk Papperstidn. 60:158
Bj6rkman, A.; Person, B. 1957b. Svensk Papperstidn. 60:285
Bland, D. E. 1961. Holzforschung 15:102
Bland, D. E.; Ho, G.; Cohen, W. E. 1950. Austral. J. Sci. Res. A 3:642
Brunow, G. 1967. Soc. Sei. Fenn. Comment. Phy. Math. 33 (9)
Brunow, G.; Miksche, G. E. 1972. Acta Chem. Scand. 26:1123
Brunow, G.; Miksche, G. E. 1976. Appl. Polym. Sci., Appl. Polym. Symp. 28:1155
Chang, H.-m.; Cowling, E. B.; Brown, W.; Adler, E.; Miksehe, G. M. 1975. Holzforschung 29:
153
C8t6, W. A.; Simson, B. W.; Timell, T. E. 1966. Svensk Papperstid. 69:547
Cousin, H.; H~rissey, H. 1908. C. R. Acad. Sci. 146:1413
Cramer, A. B.; Hunter, J. M.; H~bert, H. 1939. J. Amer. Chem. Soc. 61:509
Creighton, R. H. J.; Gibbs, R. D.; Hibbert, H. 1944. J. Amer. Chem. Soc. 6 6 : 3 2
Creighton, R. H. J.; I-h'bbert, H. 1944. J. Amer. Chem. Soc. 66:37
Erdtman, H. 1933a. Biochem. Z. 258:117

216

E. Adler

Erdtman, H. 1933b. Ann. 503:283


E~dtman, H. 1941. Svensk Papperstid. 44:243
Erdtraan, H. 1950. Research 3 : 6 3
Erdtman, H.; Leopold, B. 1948. Acta Chem. Scand. 2:535
Erdtman, H.; Leopold, B. 1949. Aeta Chem. Scand. 3:1358
Erickson, M.; Laxsson, S.; Miksche, G. E. 1973a. Acta Chem. Scand. 27:127
Eriekson, M.; Larsson, S.; M~sche, G. E. 1973b. Acta Chem. Scand. 27:903
Erickson, M.; Larsson, S.; Miksche, G. E. 1973c. Acta Chem. Scand. 27:1673
Erickson, M.; Miksche, G. E. 1972. Acta Chem. Scand. 26:3085
Erickson, M.; Miksche, G. E. 1974a. Holzforschung 28:135
Erickson, M.; Miksche, G. E. 1974b. Phytochemistry 13:2295
Erickson, M.; Miksche, G. E.; Somfai, L 1973c. Holzforschung 27:113
Erickson, M.; Miksche, G. E.; Somfai, I. 1973d. Holzforschung 27:147
Eriksson, ~.; Lindgren, B. 1977. Svensk Papperstidn. 80:59
Fengel, D. 1976. Holzforschung 30:143
Fengel, D.; Przyklenk, M. 1973. Svensk Papperstidn. 78:617
Fergus, B. J.; Procter, A. R.; Scott, J. A. N.; Goring, D. A. I. 1969. Wood Sci. Technol. 3:117
Fergus, B. J.; Goring, D. A. I. 1970a. Holzforschung 24:118
Fergus, B. J.; Goring, D. A. I. 1970b. Holzforschung 24:113
Freudenberg, K. 1939. Fortschritte der Chem. org. Naturstoffe 2 : 1
Freudenberg, K. 1966. Adv. Chem. Set. 5 9 : 1
Freudenberg, K.; Chen, C.-L. 1967. Chem. Bet. 100:3683
Freudenberg, K.; Chen, C.-L.; Cardinale, G. 1962. Chem. Ber. 95:2814
Freudenberg, K.; Chen, C.-L.; Harkin, J. M.; Nimz, H.; Rermer, H. 1965. Chem. Comm. 224
Freudenberg, K.; Diirr, W. 1933. In: Klein, G. (Ed.): I-landbuch der Pflanzenanalyse. Vol. 3.
Springer, Wien, 125
Freudenberg, K.; Eisenhut, W. 1955. Chem. Ber. 88:626
Freudenberg, IC; Engler, K.; Flickinger, E.; Sobek, A.; Klink, F. 1938. Bet. Deutsch. Chem.
Ges. 71:1810
Freudenberg, K.; Grion, G. 1959. Chem. Ber. 92:1355
Freudenberg, IC; Grion, G.; Harkin, J. M. 1958. Angew. Chem. 70:743
Freudenberg, K.; Harkin, J. M. 1960. Chem. Ber. 93:2814
Freudenberg, K.; Harkin, J. M. 1964. Holzforschung 18:166
Freudenberg, K.; Harkin, J. M.; Werner, H.-K. 1964. Chem. Bet. 97:909
Freudenberg, K.; Hiibner, H. H. 1952. Chem. Bet. 85:1181
Freudenberg, K.; Janson, A.; Knopf, E.; Haag, A. 1936. Bet. Deutsch. Chem. Ges. 69:1415
Freudenberg, K.; Jones, K.; Rermer, H. 1963. Chem. Ber. 96:1844
Freudenberg, K.; Lautsch, W.; Engler, K. 1940. Ber. Deutsch. Chem. Ges. 73:167
Freudenberg, K.; Maercker, G.; Nimz, H. 1964. Chem. Ber. 97:903
Freudenberg, K.; Neish, A. C. 1968. Constitution and Biosynthesis of Lignin. Springer, Berlin
Freudenberg, K.; Nimz, H. 1966. Chem. Commun. 132
Freudenberg, K.; Reznik, H.; Fuchs, W.; Reichert, M. 1955. Naturwissenschaften 4 2 : 2 9
Freudenberg, IC; Richtzenhain, H. 1942. Arm. 552:126
Freudenberg, K.; Richtzenhain, H. 1943. Ber. Deutsch. Chem. Ges. 76:997
Freudenberg, K.; SehliJter, H. 1955. Chem. Ber. 88:617
Freudenberg, K.; Sch~ter, H.; Eisenhut, W. 1954. Natttrwissenschaften 4 1 : 5 7 6
Freudenberg, K.; Zocher, H.; Dlixr, W. 1929. Ber. Deutsch. Chem. Ges. 62:1814
Gellerstedt, G. 1976. Svensk Papperstidn. 79:537
Gibbs, R. D. 1958. In: E. Thimann (Ed.): The Physiology of Forest Trees. Rona|d Press, New
York, p. 269
Gierer, J. 1954. Acta Chem. Scand. 8:1319
Gierer, J. 1970. Svensk Papperstidn. 73:571
Gierer, J.; Lenz, B.; Nor~n, l.; Sbderberg, S. 1964. Tappi 47:233
Gierer, J.; Lenz, B.; Wallin, N. H. 1964. Acta Chem. Scand. 18:1469
Gierer, J.; Petterson, I.; Smedman, L.-/~.; Wennerberg, I. 1973. Acta Chem. Seand. 27:2083

Lto'nin chemistry-past, present and future

217

Glasser, W. G.; Glasser, H. R.; Nimz, H. H. 1976. Macromolecules 9 : 8 8 6


H6rissey, H.; Doby, G. 1909. J. Pharm. Claim., Paris (6) 3 0 : 2 8 9
Higuehi, T. 1958. Biochemistry (Japan) 4 5 : 5 1 5
Higuchi, T.; Brown, S. A. 1963. Can. J. Biochem. Physiol. 4 1 : 6 5
Higuchi, T.; Nakatsubo, F.; Ikeda, Y. 1974. Holzforschung 2 8 : 1 8 9
Holmberg, B. 1936. Svensk Papperstidn. 3 9 : 1 1 3
Hruffiord, B. F. 1971. In: Sarkanen, K. V.; Ludwig, C. H. (Eds.): Lignins. Wiley-Interscience,
New York, Chapter 12
Johansson, B.; Miksche, G. E. 1972. Acta Chem. Stand. 2 6 : 2 8 9
Kirk, T. K.; Chang, H.-m.; Lorenz, L.F. 1975. Wood Sci. Technol. 9 : 8 1
Klason, P. 1897. Svensk Kem. Tidskr. 9 : 1 3 3
Klason, P. 1908. Bericht tiber die Hauptversammlung des Vereins der ZeUstoff- und Papierchemiker, p. 52
Koshijima, T.; Yaku, F.; Tanaka, R. 1976. J. AppL Polym. ScL, Appk Polym. Syrup. 2 8 : 1 0 2 5
Kratzl, K. 1948. Monatsh. Chem. 7 6 : 1 7 3
Kratzl, K.; BiUek, G.; Klein, E.; Buchtela, K. 1957. Monatsh. Chem. 8 8 : 7 2 1
Kratzl, K.; Kisser, W.; Graf, A.; Hofbauer, G. 1959. Monatsh. Chem. 9 0 : 5 2 6
Kringstad, K. P.; Cheng, C. W. 1969. Tappi 5 2 : 2 3 8 2
Kulka, M.; Fisher, H. E.; Baker, S. B.; Hfbbert, H. 1944. J. Amer. Chem. Soc. 6 6 : 3 9
Lai, Y. Z.; Sarkanen, K. V. 1971. In: Sarkanen, K. V.; Ludwig, C. H. (Eds.): Lignins. WileyInterscience, New York, 199
Lange, P. 1954. Svensk Papperstidn. 5 7 : 5 2 5
Larsson, S.; Miksche, G. E. 1967. Acta Chem. Scand. 2 1 : 1 9 7 0
Larsson, S.; Miksche, G. E. 1969a. Acta Chem. Scand. 2 3 : 9 1 7
Larsson, S.; Miksche, G. E. 1969b. Acta Chem. Scand. 2 3 : 3 3 3 7
Larsson, S.; Miksche, G.E. 1971a. Acta Chem. Scand. 2 5 : 6 4 7
Larsson, S.; Miksche, G. E. 1971b. Acta Chem. Scand. 2 5 : 6 7 3
Larsson, S.; Miksche, G. E. 1972. Acta Chem. Scand. 26:2031
Latif, A. M. 1968. Ph. D. Thesis, Univ. of Washington, Seattle
Lemieux, R. V.; yon Rudloff, E. 1955. Can. J. Chem. 33: 1701, 1710
Leopold, B.; Malmstr/Sm, I.-L. 1952. Acta Chem. Scand. 6 : 4 9
Lindgren, B. O. 1952. Svensk Papperstidn. 5 5 : 7 8
Lindgren, B. O.; Mikawa, H. 1957. Acta Chem. Scand. 1 1 : 8 2 6
Lindgren, B. O.; Saed~n, U. 1952. Acta Chem. Scand. 6 : 9 6 3
Lundquist, K. 1964. Acta Chem. Stand. 1 8 : 1 3 1 6
Lundquist, K. 1970. Acta Chem. Scand. 2 4 : 8 8 9
Lundquist, K. 1973a. Acta Chem. Scand. 2 7 : 2 5 9 7
Lundquist, K. 1973b. Thesis, Univ. of Technology, G6teborg
Lundquist, K. 1973c. Svensk Papperstidn. 7 6 : 7 0 4
Lundquist, K.; Ericsson, L. 1970. Aeta Chem. Scand. 24:3681
Lundquist, K.; Hextlund, K. 1967. Acta Chem. Scand. 2 1 : 1 7 5 0
Lundquist, K.; Hedlund, K. 1971. Aeta Chem. Scand. 2 5 : 2 1 9 9
Lundt uist, K.; Kirk, K. 1971. Acta Chem. Stand. 2 5 : 8 8 9
Lundc uist, K.; Lundgren, R. 1972. Acta Chem. Scand. 2 6 : 2 0 0 5
Lundc uist, K.; Miksche, G. E. 1965. Tetrahedron Lett. 2131
Lundc uist, K.; Miksche, G. E.; Ericsson, L.; Berndtson, L. 1967. Tetrahedron Lett. 4587
Lundc uist, K.; Ohlson, B.; Simonson, R. 1977. Svensk Papperstidn. 8 0 : 1 4 3
Lundc uist, K.; Olsson, T. 1977. Aeta Chem. Scand. In print
Lundc uist, K.; Remmerth, S. 1975. Acta Chem. Saand. B 2 9 : 2 7 6
Lundc uist, K.; Simonson, R. 1975. Svensk Papperstkln. 7 8 : 3 9 0
Marton, J.; Adler, E. Acta Chem. Scand. 1 5 : 3 7 0
Matron, J.; Adler, E.; Persson, K.-I. 196I. Acta Chem. Scand. 1 5 : 3 8 4
Miksche, G. E. 1972. Acta Chem. Scand. 2 6 : 4 1 3 7
Miksche, G. E. 1973. Acts Chem. Scand. 2 7 : 1 3 5 5
Miksche, G. E.; Gratzl, J.; Fried-Matzka, M. 1966. Acta Chem. Scand. 2 0 : 1 0 3 8

218

E. Adler

Miksc~e, G. E.; Yasuda, S. 1976. Ann. 1323


Musha, Y.; Goring, D. A. I. 1975. Wood Sci. Technol. 9 : 4 5
N i m z H. 1965. Chem. Bet. 9 8 : 3 1 6 0
Nimz H. 1966a. Chem. Ber. 9 9 : 4 6 9
Nimz H. 1966b. Holzforschung 2 0 : 1 0 5
Nimz H. 1966c. Angew. Chem. Int. Ed. Engl. 5 : 8 4 3
N i m z H. 1967. Chem. Bet. 1 0 0 : 1 8 1
Nimz H. 1969. Chem. Bet. 1 0 2 : 7 9 9
Nimz H. 1974. Angew. Chem. Int. Ed. Engk 1 3 : 3 1 3
Nimz H.; Das, K. 1971. Chem. Ber. 1 0 4 : 2 3 5 9
Nimz H.: Das, K.; Minemura, N. 1971. Chem. Bet. 104:1871
Nimz, H.; Gaber, H. 1965. Chem. Bet. 9 8 : 5 3 8
Nimz, H.; Moghamb, I.; IAldemann, H.-D. 1974. Makromol. Chem. 1 7 5 : 2 5 6 3
Ogiyama, IC; Kondo, T. 1966. Tetrahedron Lett. 2083
Ogiyama, K.; Kondo, T. 1967. Tetrahedron Lett. 3144
Omori, S.; Sakak~ara, A. 1971. Mokuzai Gakkaishi 1 7 : 4 6 4
Pearl, I. A. 1942. J. Amer. Chem. Soc. 6 4 : 1 4 2 9
Peppe~, J. M.; Baylis, P. E. T.; Adler, E. 1959. Can. J. Chem. 37:1241
Pew, J. C. 1957. Tappi 4 0 : 5 5 3
Pew, J. C.; Weyna, P. 1962. Tappi 4 5 : 2 4 7
Read, D. E.; Purves, C. B. 1952. J. Amer. Chem. Soc. 7 4 : 1 2 0
Sakakibara, A.; Nakayama, N. 1962. MokuzaJ Gakkaishi 8 : 1 5 7
Sakakib~ra, A.; Takeyama, H., Morohoshi, N. 1966. Holzforschung 2 0 : 4 5
Sano, Y. 1975. Mokuzai Gakkaishi 2 1 : 5 0 8
Sano, Y.; Sakakibara, A. 1970b. Mokuzai Gakkaishi 1 6 : 1 2 1
Sano, Y.; Sakakibara, A. 1970m Mokuzai Gakkaishi 1 6 : 8 1
Sarkanen, K. V. 1971. In: Sarkanen, K. V.; Ludwig, C. H. rEds.): Lignins. Wiley-Interscience,
New York, p. 151
Surkanen, K. V.; Chang, H.-M.; Allan, G. G. 1967. Tappi 50" 587
Sarkanen, IC V.; Hergert, H. L. 1971. In: Sarkanen, K. V.; Ludwig, C. H. (Eds.): Lignins. WileyInterscience, New York, Chapter 3
Sarkanen, K. V.; Ludwig, C. H. (Eds.): 1971. Lignins. Occurrence, Formation, Structure and
Reactions. Wiley-lnterscience, New York
Scott, J. A. N.; Procter, A. tL; Fergus, B. J.; Goring, D. A. I. 1969. Wood Sci. Technol. 3 : 7 3
Tiernann, F.; Mendelsohn, B. 1875. Ber. Deutsch. Chem. Ges. 8 : 1 1 3 9
TimeU, T. E. 1965. Advan. Carbohyd. Chem. 2 0 : 4 4 3
Timell, T. E. 1967. Wood Sci. Technol. 1 : 4 5
WaUis, A. F. A. 1971. In: Sarkanen, K. V.; Ludwig, C. H. (Eds.): Lignins. Wiley-Interscience,
New York, p. 3 5 5 - 3 5 6
West, E., Macinnes, A. S.; l-h'bbert, H. 1943. J. Amer. Chem. Soc. 6 5 : 1 1 8 7
WtUst/~tter, R.; Zechmeister, L. 1913. Bet. Deutsch. Chem. Ges. 4 6 : 2 4 0 5
Yamasaki, T.; Higuchi, T. 1971. Mokuzai Gakkaishi 1 7 : 1 1 7

(Received June 14, 1977)


Professor Erich Adler
Department of Organic Chemistry
Chalmets University of Technology
S-402 20 G6teborg, Sweden

S-ar putea să vă placă și