Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

11

The Chemistry of Condensed Tannins

11.1 Introduction
Condensed tannin extracts consist of flavonoid units that have undergone
varying degrees of condensation. They are invariably associated with their
immediate precursors (flavan-3-ols, flavan-3,4-diols), other flavonoid
analogs [1, 2], carbohydrates, and traces of amino and imino acids [3].
Monoflavonoids and nitrogen-containing acids are present in concentra-
tions that are too low to influence the chemical and physical characteris-
tics of the extract as a whole. However, the simple carbohydrates (hexoses,
pentoses, and disaccharides) and complex glucuronates (hydrocolloid
gums) as well as oligomers derived from hydrolyzed hemicelluloses are
often present in sufficient quantity. Equally, carbohydrate chains of various
length [4, 5] are also sometimes linked to flavonoid units in the tannin.
O

O
n
OH
OH
HO HO O
O OH
O O
O OH
OH O O
m
OH
OH O

OH

All these materials are often present in sufficient quantities to decrease


and/or increase viscosity, and excessive variation in their percentages alters

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (239–266) ©  2019
Scrivener Publishing LLC

239
240 Adhesives for Wood and Lignocellulosic Materials

the physical properties of the natural tannin extract independently of the


contribution of the degree of condensation of the tannin.
Monoflavonoids, also known as “phenolic non-tannins” (to which flavo-
noid dimers belong too), represent the most studied group in the commer-
cially important tannin extracts because of their relative simplicity. They
comprise flavan-3,4-diols, flavan-3-ols, dihydroflavonoids (flavonols), fla-
vanones, chalcones, and coumaran-3-ols, thus most of the known classes
of flavonoid analogs [1, 2, 6, 7]. Typical are those of the black mimosa bark
extract (Acacia mearnsii, formerly mollissima, de Wildt) where the four pos-
sible combinations of resorcinol and phloroglucinol (A-rings) with catechol
and pyrogallol (B-rings) coexist, although these monoflavonoids consti-
tute a minor percentage (3–5%) of the total phenolics of the tannin extract
[1]. In mimosa tannin extract, only flavan-3,4-diols and certain flavan-
3-ols (catechin) participate in tannin formation.

OH
OH
OH
8 B
HO 7 O 8 B
OH OH 7 O
A OH
6
4 A
5 OH 6
4
5 OH
profisetinidin prorobinetinidin
Mw = 272.3 Da Mw = 288.3 Da
OH
OH OH
8 B B
HO 7 O HO 7
8
O
OH OH
A A
6 4 4
6
5 OH 5 OH
OH procyanidin OH prodelphinidin
Mw = 288.3 Da Mw = 304.3 Da

In other tannins, epicatechin, delphinidin, and catechin gallate are


present, always in minor percentages but with some exception. The main
exception is cube gambier tannin extract where monoflavonoids, mainly
catechin, can constitute up to 50% of the total extract. In mimosa bark
tannin extract, each of the four combinations of resorcinol and phloro-
glucinol (A-rings) with catechol and pyrogallol (B-rings) are present. In
this tannin, the main polyphenolic pattern is represented by flavonoid
analogs based on roninetinidin, and thus on resorcinol A-ring and pyro-
gallol B-ring. This pattern is reproduced in approximately 70% of the phe-
nolic part of the tannin. The secondary but parallel pattern is based on
fisetinidin and thus on resorcinol A-rings and catechol B-rings. This rep-
resents about 25% of the total polyphenolic bark fraction. Superimposed
on these two predominant patterns are two minor groups of A- and B-ring
The Chemistry of Condensed Tannins 241

combinations. These are based on phloroglucinol (A-ring)–pyrogallol


(B-ring) (gallocatechin/delphinidin) and on phloroglucinol (A-ring)–cat-
echol (B-ring) flavonoids (catechin/epicatechin). These four patterns con-
stitute 65–84% of commercial mimosa bark extract. The remaining parts
of mimosa bark extract are the so-called “non-tannins”, this definition
coming from the leather industry where any polyphenolic oligomer higher
and comprising atrimer is considered a “tannin”. It must be pointed out
that the percentage of non-tannins varies considerably from tannin extract
to tannin extract. For example, pecan nut tannin extract, a predominantly
delphinidin tannin, contains no more than 5% of non-phenolic non-tan-
nins [8,  9]. The percentages are different even for different commercial
tannins [8, 9]. The non-phenolic non-tannins can be subdivided into car-
bohydrates, hydrocolloid gums, and some amino and imino acid fractions
[9, 10]. For example, in commercial mimosa bark extract, the carbohy-
drates 1-pinitol and sucrose predominate, with glucose in a smaller pro-
portion [9, 10]. The hydrocolloid gums contribute between 3% and 8%
to the total weight of commercial mimosa bark extract and are a major
contributor to its viscosity [9, 10]. Imino acids such as L-pipecolic acid,
L-4-hydroxy-trans-pipecolic acid, and L-proline and traces of the amino
acids arginine, alanine, aspartic acid, glutamic acid, and serine have also
been reported [9, 10].
Similar flavonoid A- and B-ring patterns, although slightly different,
have been found for the other major commercial tannin extract, quebracho
(Schinopsis lorentzii and balansae) wood tannin extract, where a predomi-
nance of fisetinidin rather than robinetinidin in the constituent flavonoid
units has been determined [11]. In quebracho extract, there is apparent
absence of quercetrin and myretrecin [6, 9, 11, 12] and much lower propor-
tion or even absence of catechin and gallocatechin (delphinidin) and thus
practically lack of or much lower level of phloroglucinol-like A-rings. This
difference becomes very important from a structural point of view, as it
will be indicated later, as while the quebracho tannins have mainly a linear
structure, mimosa tannin presents a more branched structure, with very
important effects on their viscosity, their stability, and their non-leather
use. Similar couplings of A- and B-ring types occur also in Douglas-fir and
hemlock bark tannins.
However, different A- and B-ring couplings occur in pine and other bark
tannin species [4, 9, 13–20]. In this, only two main patterns occur, pre-
dominantly phloroglucinol A-ring with catechol B-ring and two secondary
patterns in much lower proportion of phloroglucinol A-ring with phenol
B-ring (afzelechin) or of fisetinidin (resorcinol A-ring, catechol-B-ring)
[9, 13–20]. In several procyanidin tannins such as some pine tannins and
242 Adhesives for Wood and Lignocellulosic Materials

exotic African woods, catechin gallate or gallocatechin gallate also occurs


as a constitutive unit of the tannin.
OH
OH OH OH
HO O HO O HO O
OH OH OH
O O O
OH OH OH OH
O O O
OH OH OH
OH OH OH

Roux et al. [6] have shown that condensation of robinetinidin and fise-
tinidin flavonoid units is based on a condensation forming a 4,6 interflavo-
noid link, following an initial 4,8 interflavonoid link formation between a
phloroglucinol A-ring type unit, such as catechin, and a resorcinol A-ring-
type units. For a time, it was thought that the positioning of the phloroglu-
cinolic flavonoid unit was as the “lower” terminal unit of a 4,8 interflavonoid
linkage. While extracts such mimosa show oligomers of this type [9, 21, 22],
the latter finding of the so-called “angular” tannins indicated that the phlo-
roglucinolic unit was not just the lower terminal unit of an oligomer but was
a central unit to which two resorcinol A-ring-type units are linked [9, 23].
OH

HO O
OH

OH
OH

HO 8 9 O
HO 2
OH
7
6
5 10 3
4 OH

OH
O
OH Angular tannin

OH

OH
The Chemistry of Condensed Tannins 243

Moreover, an even later analysis derived from comparative applied


results on the reactivity of aldehydes with different tannins [8, 9, 24]
showed that the phloroglucinol A-ring type units, be it catechin, epicat-
echin, or gallocatechin (delphinidin), was a branching point in the oligo-
mers of some tannin type (mimosa), leading to the concept of such tannins
having been heavily “branched” rather than just “linear”.
OH

HO O
OH

OH
OH

HO 8 9 O
HO 7 2
OH

6
5 10 4 3 OH
OH
OH
O
OH HO O
OH

OH

OH

OH

This was evident as two different rates of reaction with formaldehyde


did not occur but instead viscosity increase graphs as a function of time
are smooth exponential curves indicating that only one reaction site type
is present, with the phloroglucinol unit A-ring reactive sites being blocked.
This explained the particular practical differences in the use of quebra-
cho and mimosa tannins, with quebracho tannin sometimes being subject
to a reaction of partial depolymerization while mimosa is not. Mimosa is
heavily branched while quebracho is not, and thus the latter is more easily
subject to cleavage of the interflavonoid linkage under some drastic con-
ditions [8, 9, 24].
In the case of procyanidin and prodelphinidin-type tannins, which are
composed exclusively of phloroglucinol A-ring type units, the interflavo-
noid link is always 4,8, and thus all these oligomers are “linear”.
244 Adhesives for Wood and Lignocellulosic Materials

OH

HO O
OH

OH
OH
O
HO O
OH

OH
OH
OH
HO O
OH

OH

OH

11.2 Reactions of Condensed Flavonoid Tannins


The reactions with formaldehyde and other aldehydes will not be discussed
here, but other important reactions of tannin will.

11.2.1 Hydrolysis and Acid and Alkaline Condensation


When heated in the presence of strong mineral acids, tannins are subject
to two competitive reactions. One is degradative leading to anthocyani-
dins and catechin formation as illustrated in Scheme 11.1 [6, 9], whereas
the second is condensative as a result of hydrolysis of the p-hydroxybenzyl
ether links of the flavonoid heterocyclic ring. The carbonium ions formed
condense randomly with nucleophilic centers on other tannin units to
form “phlobatannins”, “phlobaphenes”, or “tanner’s red”, insoluble and no
further usable [25].
The Chemistry of Condensed Tannins 245

OH

HO O HO OH
OH

OH OH OH
OH
HO O HO O
OH OH

OH OH
OH OH

OH
HO HO OH
OH
OH

O O OH
Phlobatannin
OH
OH

Scheme 11.1

Even more complex insoluble structures, such as phlobaphenes, occur,


of formula

HO HO OH
OH

OH
HO

HO O O
OH

OH

OH O
HO

OH

OH
246 Adhesives for Wood and Lignocellulosic Materials

A system to stop the condensative reaction to go to phlobaphenes has


been found and used industrially [25], to greatly improve the percentage
extraction yield of tannins. This method consists in using a simple mole-
cule such as urea to block the carbonium ion formed.

OH

HO O HO OH
OH
OH OH
OH OH
HO O OH HO O
OH

OH OH
OH OH

NH2CONH2

H2NCONH
OH

HO OH OH
OH
OH
HO O
OH

OH
OH

Different tannins behave differently as well. For example, procyan-


idin tannins (phloroglucinol A-ring and catechol B-ring) such as pine
bark tannins favor cleavage of the interflavonoid linkage [8, 26], with the
oligomers linearity contributing to such a behavior. Predominantly pro-
delphinidin tannins, such as pecan nut tannin, also show cleavage of the
interflavonoid bond, but due to the effect of the pyrogallol B-ring, they
can also present easy cleavage and opening of the heterocyclic pyran ring
[8, 27, 28], under certain conditions up to markedly limiting interfla-
vonoid link cleavage. Predominantly profisetinidins, such as quebracho
wood tannin, tend to behave as prodelphinidins, again confirming that
interflavonoid link cleavage is facilitated by the “linearity” of the struc-
ture of the oligomers of these tannins. For mimosa tannin, where a cer-
tain degree of branching exists, interflavonoid link cleavage practically
does not occur, with cleavage and opening of the heterocyclic pyran ring
by far being the favorite reaction.
The Chemistry of Condensed Tannins 247

11.2.2 Sulphitation
Sulphitation of tannins is one of the oldest and most useful reactions
in condensed tannins chemistry. It can be useful or damaging if car-
ried out to an excessive extent, according to the final use to which the
tannin extract is destined [9, 29]. Sulphitation allows the preparation
of tannins of lower viscosity and increased solubility and thus easier to
handle.

OH

HO O - HO OH OH
OH SO3- OH

OH OH
SO3Na

Such effects are due to:

(i) The elimination of the heterocyclic ring, which is water


repellent. It must be remembered that tannin extract water
solutions are not true solutions but hydrocolloid suspen-
sions, in which part of the tannin molecule keeps the tan-
nin in solution while other parts tend to push the tannin
out of solution [9, 30].
(ii) The introduction of a sulphonic group and a further
hydroxygroup, both hydrophilic.
(iii) The decrease in polymer rigidity, steric hindrance, and
intermolecular hydrogen bonding obtained by the open-
ing of the heterocyclic ring.
(iv) Hydrolysis of the hydrocolloid gums [9, 31] and of the
interflavonoid bond (for some tannins) [9].
(v) The increase of reactivity towards aldehydes by the trans-
formation of the ether linkage of the heterocyclic ring
into a –OH linked on to the A-rings.

While the above are considerable advantages, excessive sulphitation


can be a distinct disadvantage for some applications. Thus, the introduc-
tion of an excessive proportion of sulphonate groups promote excessive
sensitivity to moisture and water and thus deterioration of resins, plastics
and adhesive bonds that are supposed to be moisture, water, or weather
resistant [9, 32].
248 Adhesives for Wood and Lignocellulosic Materials

11.2.3 Catechinic Acid Rearrangement


Procyanidin- and delphinidin-type tannins, but also other flavonoid tan-
nins, can be subject to a reaction of rearrangement to catechinic acid. This
reaction leads to the deactivation of several reactive sites of the tannin and
thus it is to be avoided. Some manipulation errors in tannin extraction
or in drying generally cause this rearrangement. The appearance of the
catechinic acid derived directly from a flavonoid monomer and the one
derived from a dimer are as follows:

HO OH
HO O OH

OH

OH HO O
O
OH

OH
OH
O
OH

OH
OH

While the catechinic acid rearrangement is easily shown to occur in


model compounds where the reaction is carried out in solution [9, 33], it
is much less evident and more easily avoidable in tannin extracts where
the colloidal nature of the extract limits markedly its occurrence. This is
fortunate as otherwise some fast-reacting tannins such as pine, pecan,
cube gambier, etc. could not be used to produce resins, adhesives, and
other thermosetting plastics as instead they have been successfully used
for [27, 34–38]. It can be easily determined if a tannin extract is affected
by this rearrangement both by its loss of reactivity and by 13C NMR anal-
ysis [39].

11.2.4 Catalytic Tannin Autocondensation


Polyflavonoid tannins have been found to autocondense and harden when
in the presence of particular compounds acting as catalysts. Foremost is
the catalytic effect of small amounts (2–3%) of silica smoke, or nanosilica
or silicates at high pH [40]. This reaction is rapid and is markedly exo-
thermic, a concentrated solution of tannin at 40–50% in water gelling and
The Chemistry of Condensed Tannins 249

hardening at pH 12 and 25°C in 20–30 min. The strong exothermicity of


the reaction leads to this result as the temperature increases several tens of
degrees in a short period [40].

Si(OH)4
Si(OH)4 O
OH
B HO O
HO O OH
OH
OH
OH
OH
OH

OH
HO OH OH Si(OH)4
OH O
HO O
OH OH
OH HO
OH
OH -δ OH
HO OH OH OH
OH

O
OH

OH
OH

Small amounts of boric acid and AlCl3 were found to have the same
effect [40] but are much less exothermic. Interestingly, even the presence
of lignocellulosic material, such as placing the tannin on the wood sur-
face, has a catalytic effect on tannin autocondensation [41–43]. Other
coreactants seem to further favor the reaction [41–43]. This mechanism
of tannin autocondensation appears to depend on the Lewis acid behav-
ior of the reagents. It involves Lewis acid acceptance of electrons from
the ether oxygen of the flavonoid heterocyclic pyran ring with facili-
tation of base-induced pyran ring opening. The reactive C2 site cre-
ated by the heterocyclic pyran ring opening proceeds to condense with
the reactive A-ring of a flavonoid unit on another chain. This denies
to the flavonoid any possibility of rearrangement to catechinic acid or
phlobaphenes. In the SiO2 catalysis, Si has been shown to go through
a coordination state of 5. The portion of Si that has not been able to
complete the reaction due to premature hardening remains attached to
the flavonoids, in this coordination state, in the hardened network. The
Si portion that has completed the reaction and caused the hardening
reverts instead to SiO2 and is detached from the flavonoid. It restarts
the cycle to lead to complete hardening up to when diffusional prob-
lems do not stop the reaction. The reaction is rather exothermic, with
250 Adhesives for Wood and Lignocellulosic Materials

the marked increase in temperature observed contributing to the self-


acceleration of the reaction. The SiO2, boric acid, and AlCl3 catalyses
of tannin autocondensation also reverse the relative ease of cleavage
between interflavonoid bond and heterocyclic ring opening in mainly
procyanidin- and prodelphinidin-type tannins. This reaction has been
used to good effect for no-aldehyde interior-grade tannin adhesives for
wood panels [44]. As the cross-linking obtained is only acceptable for
interior joints, coupling with traditional polycondensation reactions
can be implemented easily [42, 43].

11.2.5 Tannin Complexation of Metals


Tannins readily complex metal ions [45]. This characteristic is at the base
of a number of industrial applications. Thus, this characteristic is used
to capture or precipitate toxic metals in water [46, 47], to isolate a rare
metal such as Germanium from the copper matrix where it is mined,
for paint primers for metal application, and for several other applica-
tions. An old example is the formation of Fe complexes, used to prepare
intensely black/violet inks by formation of ferric tannates. These coordi-
nation complexes are due to the ortho-diphenol hydroxyl groups on the
tannin B-rings.

O O
III
Fe
HO O O OH
O O

OH HO

11.2.6 Tridimensional Structure


While there is an abundant literature in chemical journals on the tridi-
mensional structure of flavonoid monomers, one point in which only scant
literature exists is on the three-dimensional spatial configuration of flavo-
noid oligomers. Only one molecular mechanics study on this subject exists
[48]. This study shows the correlation that exists between the applicability
of these materials and their 3D structure. For example, a tetraflavonoid of
4,8-linked catechins, all 3,4-cis, is in helix configuration, and when looked
along the helix axis, a characteristic structure presenting all the 4 B-rings
pointing outwards appears.
The Chemistry of Condensed Tannins 251

Such a structure, rendering particularly available the hydroxyl groups of


the B-rings obviously facilitates their use and reactions, such as adhesion
to a lignocellulosic substrate, formation of metallic coordination complexes
[45–47], formation of polyurethanes with and without isocyanates [49, 50],
and others where reaction of the B-ring is of interest [51], such as cross-
linking at pH 10 and higher. Conversely, the “spring-like” structure con-
tributes to some of the “resuscitation” behaviors of some plants by holding
together the cellular walls and avoiding cellular walls cracking on drying [52].

11.2.7 Reactivity and Orientation of Electrophilic


Substitutions of Flavonoids
The relative accessibility and reactivity of flavonoid units is of interest for
their use in resins and adhesives. The C8 site on the A-ring is the first one
to react, for example, with an aldehyde, and is the site, when free of high-
est reactivity [6, 8]. The C6 site on the A-ring is also very reactive but less
than the C8 site as this latter presents lower steric hindrance too [6, 8]. The
reactions involve in general only these two sites on the A-ring. The B-ring
is particularly unreactive. A low degree of substitution at the 6 site of the
B-ring can occur. In general, at higher pHs, such as pH 10, the B-ring starts
to react too, contributing to cross-linking as well [51, 53].
1
3 2 3
OH OH
6' 6'
HO 8 O HO O
OH 8 OH
6 6
OH OH
2 OH 1
252 Adhesives for Wood and Lignocellulosic Materials

Thus, for catechins and phloroglucinol A-ring type flavonoids, the reac-
tivity sequence of the sites is C8 > C6 > C6 when these are free. For rob-
inetinidin and fisetinidin and thus for resorcinol A-ring-type flavonoids,
the reactivity sequence is modified to C6 > C8 > C6 due to the greater
accessibility and lower possibility of steric hindrance of the C6 site [6, 9].
The curve of gel time of flavonoid tannins with aldehydes has always the
shape of a bell curve. The longer gel time is at around pH 4 and the fastest
gel times are at lower pHs and higher pHs. The curve reaches an almost
asymptotic plateau of very high reactivity and short gel time at around pH
10 and higher and a fast reactivity too at pHs lower than 1–2 [8, 9]. The
shape of this curve is always the same, but the gel time value is different
for different tannins, being slower for mimosa and quebracho, and much
faster for procyanidin-type tannins (such as pine) [54, 55].

11.2.8 Influence of Tannin Colloidal Behavior on Reactions


Water solutions of 40–50% polyflavonoid tannin extracts appear to be in a
colloidal state as indicated by their zeta-potentials [55]. This is caused by
both the presence of noticeable proportions of hydrocolloid gums (frag-
ments of hemicelluloses) as well as the presence of higher-molecular-mass
tannins. 13C NMR has confirmed that, during chemical treatment of the
tannin extracts, reactions occur in these colloidal solutions that would
not be likely to occur in noncolloidal solutions such as used in model
compound experiments. These reactions center on reactions occurring in
the part of the tannin that is the non-aqueous environment within the
colloidal micelles, away from water, within which reagents can migrate.
An example of this is the role of an organic anhydride by addition of ace-
tic or maleic anhydride to hot, concentrated water solutions of tannin, a
reaction used to increase the reactivity of the tannin [55, 56]. While part
of the anhydride is hydrolyzed to acid in water, part of it does instead react
within the colloidal micelles to give acetylation and maleation of some
flavonoids of the tannin [55, 56] contributing to the marked improvement
of reactivity towards aldehydes of the tannin by allowing an alpha-set
approach [57]. Such reactions appear to be particularly beneficial to the
quebracho and mimosa tannin extracts and have some noticeable positive
effects on the higher-reactivity procyanidin-type tannin extracts, such
as pine bark tannin, but less due to its already much higher reactivity.
However, they may have deleterious effects on higher-reactivity tannin
extracts such as the pecan nut prodelphinidin tannin. The reason for the
latter behavior is the very low level, near absence of colloidal gum in the
extract, and thus very low level of colloidal state, if any. These results are in
The Chemistry of Condensed Tannins 253

line with the established zeta-potentials of the different tannin extracts


measured [55, 56].

11.2.9 New and Unusual Tannin Reactions


Recently, a number of reactions of tannins that could be useful for a
number of different applications have come to light. The first of these is
the reaction of flavonoid tannins with concentrated aqueous ammonia
[58]. Catechin was also used as a model compound and treated under the
same conditions mimosa tannin extract was treated. Solid-state 13C NMR
and matrix-assisted laser desorption/ionization time of flight (MALDI-
TOF) spectroscopy showed that, unlike what was recently thought [59],
amination is not always regioselective and leads to the conversion of one
single –OH group in C4 into a –NH2 group. New reactions have been
evidenced, clearly leading to multiamination of several phenolic hydrox-
ygroups, heterocycle opening, and oligomerization and cross-linking
through the formation of –N= bridges between flavonoid units, as shown
here:
OH OH
HO HO N OH
OH
OH
HO
OH
OH 589Da

The amination reaction of condensed tannins was used, among


others, to totally eliminate synthetic materials in the preparation of
non-isocyanate polyurethanes derived from tannins [60].
Follow-up of the reaction with ammonia was the development of rapid
cross-linking by reaction of tannin extract with diamines and polyamines
[61]. Reaction of a condensed flavonoid tannin, namely, mimosa tannin
extract with a hexamethylene diamine, has been investigated. Catechin
was also used as a flavonoid model compound and treated under similar
conditions. Solid-state CP-MAS 13C NMR and MALDI-TOF mass spec-
troscopy showed that polycondensation compounds leading to resins
were obtained by the reaction of the amines with the phenolic hydroxy
groups of the tannin. Simultaneously, a second reaction leading to the
formation of ionic bonds between the two groups occurred. These new
reactions have been shown to clearly lead to the reaction of several phe-
nolic hydroxyl groups, and flavonoid unit oligomerization, to form hard-
ened resins. MALDI-TOF analysis allowed us to observe the presence of
compounds of the type
254 Adhesives for Wood and Lignocellulosic Materials

HO

OH O OH
HO
HO O NH
NH HO OH
O NH
OH HO NH2
OH
OH
HO H
HO O
NH O
Na
NH2
OH
OH

Clearly indicating how polymerization and cross-linking occurred.


The third very novel reaction is based on the reaction, oligomeriza-
tion, and cross-linking of tannins by triethyl phosphate (TEP) [62] in
the presence or absence of ammonia (this latter being preferable to the
yield). Reaction of condensation and cross-linking of catechin monomer
as a model of condensed (flavonoid) tannin extracts and of mimosa tan-
nin itself, as well as of resorcinol with TEP, have been investigated. Solid-
state CP-MAS 13C NMR, 31P NMR, and MALDI-TOF spectroscopy studies
revealed that reaction occurs mainly on the C3 of the flavonoid heterocycle
ring and on the aromatic C4 and C5 carbons of the flavonoid B-ring,
while TEP does not appear to react on the A-ring. Structures of the type

O O
P
O O

HO O H
O

O
O OH
P
OH O
O

O OH

O
O
P
O OH
O
The Chemistry of Condensed Tannins 255

were obtained and the tannin was cross-linked, with or without being
first reacted with ammonia. The resin so obtained can produce hard
thermoset plastics and films resistant up to temperatures in excess of
400°C.
A difference in the relative proportions of these two reaction sites
for tannin and catechin has been noticed. The main reaction for the
tannin appears to occur on the C3 site of the heterocycle ring while cat-
echin monomer reacts principally on the OH of B-ring. This difference
could be explained by the lower mobility of the tannin, due to its higher
molecular weight and to its colloidal state. The reactions appear to be
dependent on the temperature. The reaction appears to have a tem-
perature of activation below which it does not appear to occur. Thus, it
occurs readily at 185°C but does not at 100°C. This aspect needs further
investigation. FTIR analysis confirmed the results of the MALDI-TOF
and NMR analysis. According to thermogravimetric analysis, materials
obtained from the reaction of tannin with TEP showed high thermal
stability. In this context, the potential of this reaction has been evalu-
ated for the production of new heat-resistant biomaterials and lacquers
[62–64].

11.2.10 Modern Instrumental Methods of Analysis


Analysis of tannin features are done by a number of traditional meth-
ods of all types. However, two instrumental methods have particularly
distinguished themselves in identifying structural features and oligomer
type, size, and distribution to better understand these materials. These
are 13C NMR, both in liquid phase or solid state, and MALDI-TOF mass
spectroscopy. These two techniques are particularly useful not only in
the analysis of tannins but also in the analysis of their reaction prod-
ucts and of the resins derived from them. Below are given the charac-
teristic 13C NMR shifts of a catechin monomer and a range of tannin
extracts from which the structural features of any condensed tannin can
be deduced (Table 11.1) [8].
Equally, MALDI-TOF techniques have allowed the determination
of oligomer type, molecular weight, and distribution of a number of
tannin extracts and products obtained from their reactions [14–20,
65–71]. From a MALDI-TOF spectrum, one can distinguish mono-
mers, dimers, trimers, higher oligomers, and so on. Figure 11.1 shows
MALDI-TOF spectra of a commercial mimosa tannin extract and Table
11.2 shows the more interesting oligomer assignments derived by such
spectra.
256

Table 11.1 Comparative 13C NMR shift assignments and relative band intensities (%) for pure catechin and five types of polyflavonoid
tannin extracts.
Assignment (ppm)

Phloro-
glucinol
inter- Catechinic
C5, C7 C9 C3', C4' C1 C6'* C5', C2'* flavonoid C10 C6 C8 C2 C3 acid C4
(156–157) (155) (145–146) (131) (120–121) (115–117) (110) (101) (96–98) (95–96) (81–82) (67–68) (31–32) (27–28)

Pure 97, 100 58 72, 76 73 97 98, 96 – 68 88 97 92 93 – 80


catechin

Mimosa 53 30 100 44 25 26 51 19 21 31 – – – 21
bark

Quebracho 63 33 100 66 56 99 40, 52 25 20 20 – – – 32


wood

Pine bark 71 71 100 49 39 76 60 37 47 23 – – – –

Pecan nut 66 70 100 62 30 33 66 42 40 29 – – 79 37


pith

Gambier 100 50 100 55 69 106, 95 27 50 60 28 – – 61 61


Adhesives for Wood and Lignocellulosic Materials

Source: Ref. [8].


The Chemistry of Condensed Tannins 257

100 375

90

80 890
70

60 551
% int.

50 1179
40

30 1467

20
488 727 1756
10 988 2045
2333
0
400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400 3600 3800 4000
Mass/Charge
(a)

890
100
906 305 Da
90
80 289 Da
70
60 1179
273 Da
% int.

50 1195
874
40 602 1163

30
922
20 664 727
634 700 750 781 814 858 1147 1211
10 687 759 840 943 988

0
600 650 700 750 800 850 900 950 1000 1050 1100 1150 1200 1250 1300
Mass/Charge
(b)

Figure 11.1 MALDI mass spectrum of (a) natural mimosa tannin extract. (b) Details of
the 600–1300 Da range with indication of the relevant 288-Da repeat unit [66].

11.3 Conclusions
The chemistry and characteristic reactions of condensed flavonoid tan-
nins have been the basis for their extended industrial utilization. It is
on the basis of this chemistry that many heavily or totally biosourced
materials have been developed. Among these are industrialized wood
panel adhesives [31, 34, 37, 72], industrialized wood laminating adhe-
sives [73–75], fire-resistant biosourced rigid and flexible foams [76–87],
foams for acoustic isolation [88], hard plastics [89], grinding disks for
angle grinders [90], automotive brake pads [91], paper impregnating res-
ins [92–94], high-pressure laminates [95], impregnated fiber composites
258 Adhesives for Wood and Lignocellulosic Materials

Table 11.2 MALDI fragmentation peaks for industrial mimosa tannin extract
of Figure 11.1. Note that the predominant repeat units in this tannin is 288 Da,
indicating that this tannin is predominantly a prorobinetinidin [66].
M + Na+ M + Na+ Unit type (MW)
(exp.) (calc.) 274 290 306
Dimers
602 601 – 2 –
Trimers
858 857 2 1 –
874 873 1 2 –
or 2 – 1 Angular tannin
*890 889 1 1 1
or 1 – 2
*906 905 – 2 1 Angular tannin
or 1 – 2 Angular tannin
922 921 – 1 2 a “diangular” structure
Tetramers
1147 1145 2 2 –
or 3 – 1
1163 1161 1 3 –
or 2 1 1
*1179 1177 – 4 –
or 1 2 1
or 2 – 2
1195 1193 – 3 1 Angular tannin
or 1 1 2
1211 1209 – 2 2 Angular tannin
(Continued)
The Chemistry of Condensed Tannins 259

Table 11.2 MALDI fragmentation peaks for industrial mimosa tannin extract
of Figure 11.1. Note that the predominant repeat units in this tannin is 288 Da,
indicating that this tannin is predominantly a prorobinetinidin [66]. (Continued)
M + Na+ M + Na+ Unit type (MW)
(exp.) (calc.) 274 290 306
or 1 – 3 a “diangular” structure
Pentamers
1467 1465
Hexamers
1756 1753
Heptamers
2045 2041
Octamers
2333 2329
*Dominant fragment.

[96, 97], biosourced wood preservatives [98, 99], non-isocyanate poly-


urethane surface finishes and resins [49, 100, 101], and others such as
medical/pharmaceutical applications [102–105] (Pizzi 2008), antioxi-
dant applications [106], and precipitation of heavy metals from wastewa-
ters [46, 47]. There is hope that on the basis of the same chemistry, many
other industrial products of progressively higher added value may also be
developed in the future.

References
1. Drewes, S.E. and Roux, D.G., Condensed tannins. 15. Interrelationships of
flavonoid components in wattle-bark extract. Biochem. J., 87, 167–172, 1963.
2. Roux, D.G. and Paulus, E., Condensed tannins. 8. The isolation and distribu-
tion of interrelated heartwood components of Schinopsis spp. Biochem. J., 78,
785–789, 1961.
3. Saayman, H.M. and Roux, D.G., The origins of tannins and flavonoids in
black-wattle barks and heartwoods, and their associated ‘non-tannin’ com-
ponents. Biochem. J., 97, 794–801, 1965.
260 Adhesives for Wood and Lignocellulosic Materials

4. Drovou, S., Pizzi, A., Lacoste, C., Zhang, J., Abdalla, S., Al-Marzouki, F.M.,
Flavonoid tannins linked to long carbohydrate chains—MALDI TOF anal-
ysis of the tannin extract of the African locust bean. Ind. Crops Prod., 67,
25–32, 2015.
5. Abdalla, S., Pizzi, A., Ayed, N., Chrrier, F., Bahabry, F., Ganash, A., MALDI-
TOF and 13C NMR analysis of Tunisian Zyzyphus jubjuba root bark tannins.
Ind. Crops Prod., 59, 277–281, 2014.
6. Roux, D.G., Ferreira, D., Hundt, H.K.L., Malan, E., Structure, stereochemis-
try, and reactivity of natural condensed tannins as basis for their extended
industrial application. Appl. Polym. Symp., 28, 335–353, 1975.
7. Roux, D.G., Recent advances in the chemistry and chemical utilization of the
natural condensed tannins. Phytochemistry, 11, 1219–1230, 1972.
8. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York,
1994.
9. Pizzi, A., Tannin based adhesives, in: Wood Adhesives Chemistry and
Technology, A. Pizzi (Ed.), Marcel Dekker, New York, 1983.
10. Roux, D.G., Modern Applications of Mimosa Extract, pp. 34–41, Leather
Industries Research Institute, Grahamstown, South Africa, 1965.
11. Mitsunaga, T., Doi, T., Kondo, Y., Abe, I., Color development of proantho-
cyanidins in vanillin–hydrochloric acid reaction. J. Wood Sci., 44, 125–130,
1998.
12. Clark-Lewis, J.W. and Roux, D.G., Natural occurrence of enantiomorphous
leucoanthocyanidian: (+)-mollisacacidin (gleditsin) and quebracho(–)-
leucofisetinidin. J. Chem. Soc., 0, 1402–1406, 1959.
13. Porter, L.J., Extractives of Pinus radiata bark. 2. Procyanidin constituents.
N. Z. J. Sci., 17, 213, 1974.
14. Navarrete, P., Pizzi, A., Pasch, H., Rode, K., Delmotte, L., MALDI-TOF and
13
C NMR characterisation of maritime pine industrial tannin extract. Ind.
Crops Prod., 32, 105–110, 2010.
15. Ucar, M.M., Ucar, G., Pizzi, A., Gonultas, O., Characterisation of Pinus brutia
bark tannin by MALDI-TOF and 13C NMR. Ind. Crops Prod., 49, 679–704,
2013.
16. Abdalla, S., Pizzi, A., Ayed, N., Charrier-El-Bouthoury, F., Charrier, B.,
Bahabry, F., Ganash, A., MALDI-TOF analysis of Aleppo pine (Pinus halep-
pensis) bark tannin. Bioresources, 9, 3396–3406, 2014.
17. Abdalla, S., Pizzi, A., Ayed, N., Charrier, B., Bahabry, F., Ganash, A., MALDI-
TOF and 13C NMR analysis of Tunisian Zyzyphus jubjuba root bark tannins.
Ind. Crops Prod., 59, 277–281, 2014.
18. Saad, H., Charrier-El-Bouthoury, F., Pizzi, A., Rode, K., Charrier, B., Ayed, N.,
Characterization of pomegranate peel tannin extractives. Ind. Crops Prod.,
40, 239–246, 2012.
19. Navarrete, P., Pizzi, A., Pasch, H., Rode, K., Delmotte, L., Characterisation
of two maritime pine tannins as wood adhesives. J. Adh. Sci. Technol., 27,
2462–2479, 2013.
The Chemistry of Condensed Tannins 261

20. Vazquez, G., Pizzi, A., Freire, M.S., Santos, J., Antorrena, G., Gonzalez-
Alvarez, J., MALDI-TOF, HPLC-ESI-TOF and 13C NMR characterisation of
chestnut (Castanea sativa) shell tannins. Wood Sci. Technol., 47, 523–535,
2013.
21. Roux, D.G., Recent advances in the chemistry and chemical utilization of the
natural condensed tannins. Phytochemistry, 11, 1219–1230, 1972.
22. Hundt, H.K.L. and Roux, D.G., Condensed tannins: Determination of the
point of linkage in ‘terminal’(+)-catechin units and degradative bromination
of 4-flavanylflavan-3,4-diols. J. Chem. Soc., Chem. Comm., 0, 696–698, 1978.
23. Botha, J.J., Ferreira, D., Roux, D.G., Condensed tannins. Circular dichroism
method of assessing the absolute configuration at C-4 of 4-arylflavan-3-ols,
and stereochemistry of their formation from flavan-3,4-diols. J. Chem. Soc.
Chem. Comm., 0, 698–700, 1978.
24. Pizzi, A., Natural phenolic adhesives 1: Tannin, in: Handbook of Adhesive
Technology, 2nd edn, A. Pizzi and K.L. Mittal (Eds.), pp. 573–598, Marcel
Dekker, New York, 2003.
25. Sealy-Fisher, V.J. and Pizzi, A., Increased pine tannins extraction and wood
adhesives development by phlobaphenes minimization. Holz Roh Werkstoff,
50, 212–220, 1992.
26. Hemingway, R.M., Laks, P.E., McGraw, G.W., Kreibich, R.E., Wood Adhesives
in 1985: Status and Needs, Forest Products Society, Madison, Wisconsin,
1986.
27. Pizzi, A. and Stephanou, A., Comparative and differential behaviour of pine
vs. pecan nut tannin adhesives for particleboard. Holzforsch. Holzverwert.,
45, 2, 30–33, 1993.
28. McGraw, G.W., Rials, T.G., Steynberg, J.P., Hemingway, R.W., Plant
Polyphenols, R.W. Hemingway and P.E. Laks (Eds.), Plenum Press, New York,
1992.
29. Richtzenhain, H. and Alfredsson, B., Uber Ligninmodellsubstanzen. Chem.
Ber., 89, 378, 1956.
30. Roux, D.G., Wattle Tannin and Mimosa Extract, Leather Industries Research
Institute, Grahamstown, South Africa, 1965.
31. Pizzi, A., Wattle-based adhesives for exterior grade particleboard. For.
Prod. J., 28, 12, 42–47, 1978.
32. Pizzi, A., Sulphited tannins for exterior wood adhesives. Colloid Polym. Sci.,
257, 37–40, 1979.
33. Ohara, S. and Hemingway, R.W., Condensed tannins: The formation of
a  diarylpropanol-catechinic acid dimer from base-catalyzed reactions of
(+)-catechin. J. Wood Chem. Technol., 11, 195–208, 1991.
34. Pizzi, A., Pine tannin adhesives for particleboard. Holz Roh Werkst., 40, 293,
1982.
35. Pizzi, A., Von Leyser, E.P., Valenzuela, J., Clark, J.G., The chemistry and devel-
opment of pine tannin adhesives for exterior particleboard. Holzforschung,
47, 164–172, 1993.
262 Adhesives for Wood and Lignocellulosic Materials

36. Pizzi, A. and Stephanou, A., Fast vs. slow-reacting non-modified tannins
extracts for exterior particleboard adhesives. Holz Roh Werkst., 52, 218–222,
1994.
37. Valenzuela, J., Von Leyser, E., Pizzi, A., Westermeyer, C., Gorrini, B.,
Industrial production of pine tannin-bonded particleboard and MDF. Eur. J.
Wood Prod., 70, 735–740, 2012.
38. Pizzi, A., Valenzuela, J., Westermeyer, C., Low-formaldehyde emission, fast
pressing, pine and pecan tannin adhesives for exterior particleboard. Holz
Roh Werkst., 52, 311–315, 1994.
39. Navarrete, P., Pizzi, A., Bertaud, F., Rigolet, S., Condensed tannin reactiv-
ity inhibition by internal rearrangements: Detection by CP-MAS 13C NMR.
Maderas, 13, 1, 59–68, 2011.
40. Meikleham, N., Pizzi, A., Stephanou, A., Induced accelerated autocondensa-
tion of polyflavonoid tannins for phenolic polycondensates, Part 1: 13C NMR,
29
Si NMR, X-ray and polarimetry studies and mechanism. J. Appl. Polym.
Sci., 54, 1827–1845, 1994.
41. Garcia, R., Pizzi, A., Merlin, A., Ionic polycondensation effects on the radical
autocondensation of polyflavonoid tannins—An ESR study. J. Appl. Polym.
Sci., 65, 2623–2632, 1997.
42. Garcia, R. and Pizzi, A., Polycondensation and autocondensation networks
in polyflavonoid tannins, Part 1: Final networks. J. Appl. Polym. Sci., 70,
1083–1091, 1998.
43. Garcia, R. and Pizzi, A., Polycondensation and autocondensation networks
in polyflavonoid tannins, Part 2: Polycondensation vs. autocondensation.
J. Appl. Polym. Sci., 70, 1093–1110, 1998.
44. Pizzi, A., Meikleham, N., Dombo, B., Roll, W., Autocondensation-based,
zero-emission, tannin adhesives for particleboard. Holz Roh Werkst., 53,
201–204, 1995.
45. Slabbert, N., Complexation of condensed tannins with metal ions, in:
Plant Polyphenols: Biogenesis, Chemical Properties, and Significance, R.W.
Hemingway and P.E. Laks (Eds.), Plenum Press, New York, 1992.
46. Tondi, G., Oo, C.W., Pizzi, A., Trosa, A., Thevenon, M.-F., Metal adsorption
of tannin-based rigid foams. Ind. Crops Prod., 29, 336–340, 2009.
47. Oo, C.W., Kassim, M.J., Pizzi, A., Characterization and performance of
Rhizophora apiculata mangrove polyflavonoid tannins in the adsorption of
copper (II) and lead (II). Ind. Crops Prod., 30, 152–161, 2009.
48. Pizzi, A., Cameron, F.A., Eaton, N.J., The tridimensional structure of poly-
flavonoids by conformational analysis. J. Macromol. Sci. Chem., A23, 4, 515–
538, 1986.
49. Thebault, M., Pizzi, A., Essawy, H., Baroum, A., Van Assche, G., Isocyanate
free condensed tannin-based polyurethanes. Eur. Polym. J., 67, 513–523,
2015.
50. Pizzi, A., Tannin-based polyurethane adhesives. J. Appl. Polym. Sci., 23,
1889–1990, 1979.
The Chemistry of Condensed Tannins 263

51. Pizzi, A., Tannin formaldehyde exterior wood adhesives through flavonoid
B-ring cross-linking. J. Appl. Polym. Sci., 22, 2397–2399, 1978.
52. Pizzi, A. and Cameron, F.A., Flavonoid tannins—Structural wood compo-
nents for draught resistance mechanism of plants. Wood Sci. Technol., 20,
119–124, 1986.
53. Hillis, W.E. and Urbach, G., The reaction of (+)-catechin with formaldehyde.
J. Chem. Technol. Biotechnol., 9, 9, 474–482, 1959.
54. Hillis, W.E. and Urbach, G., Reaction of polyphenols with formaldehyde.
J. Chem. Technol. Biotechnol., 9, 12, 665–673, 1959.
55. Pizzi, A. and Stephanou, A., A 13C NMR study of polyflavonoid tannin adhe-
sives intermediates, Part 2: Colloidal state reactions. J. Appl. Polym. Sci., 51,
2125–2130, 1994.
56. Pizzi, A. and Stephanou, A., Theory and practice of non-fortified tannin
adhesives for particleboard. Holzforsch. Holzverwert., 44, 4, 62–68, 1992.
57. Pizzi, A. and Stephanou, A., On the chemistry, behaviour and cure accelera-
tion of phenol–formaldehyde resins under very alkaline conditions. J. Appl.
Polym. Sci., 49, 2157–2160, 1993.
58. Braghiroli, F., Fierro, V., Pizzi, A., Rode, K., Radke, W., Delmmotte, L.,
Parmentier, J., Celzard, A., Condensation reaction of flavonoid tannins with
ammonia. Ind. Crops Prod., 44, 330–335, 2013.
59. Hashida, K., Makino, R., Ohara, S., Amination of pyrogallol nucleus of
condensed tannins and related polyphenols by ammonia water treatment.
Holzforschung, 63, 319–326, 2009.
60. Thebault, M., Pizzi, A., Santiago-Medina, F.J., Al-Marzouki, F.M., Abdalla, S.,
Isocyanate-free polyurethanes by coreaction of condensed tannins with ami-
nated tannins. J. Renew. Mat., 5, 21–29, 2017.
61. Santiago-Medina, F., Pizzi, A., Basso, M.C., Delmotte, L., Celzard, A.,
Polycondensation resins by flavonoid tannins reaction with amines. Polymers,
9, 2, 37, 1–16, 2017.
62. Basso, M.C., Pizzi, A., Polesel-Maris, J., Delmotte, L., Colin, B., Rogaume, Y.,
MALDI-TOF and 13C NMR analysis of the cross-linking reaction of con-
densed tannins by triethyl phosphate. Ind. Crops Prod., 95, 621–631, 2017.
63. Basso, M.C., Lacoste, C., Pizzi, A., Fredon, E., Delmotte, L., Flexible tannin–
furanic films and lacquers. Ind. Crops Prod., 61, 352–360, 2014.
64. J. Polesel-Maris and I. Jutang, Antiadhesives coatings based on condensed
tannins, patent WO2017/037393 A1, assigned to SEB Development, 2017.
65. Pizzi, A. and Stephanou, A., A comparative 13C NMR study of polyflavonoid
tannin extracts for phenolic polycondensates. J. Appl. Polym. Sci., 50, 2105–
2113, 1993b.
66. Pasch, H., Pizzi, A., Rode, K., MALDI-TOF mass spectrometry of polyflavo-
noid tannins. Polymer, 42, 18, 7531–7539, 2001.
67. Radebe, N., Rode, K., Pizzi, A., Pasch, H., Microstructure elucidation of poly-
flavonoid tannins by MALDI-TOF-CID. J. Appl. Polym. Sci., 127, 1937–1950,
2013.
264 Adhesives for Wood and Lignocellulosic Materials

68. Konai, N., Raidandi, D., Pizzi, A., Meva’a, L., Characterisation of Ficus syco-
morus using ATR-FTMIR, MALDI-TOF MS and 13C NMR methods. Eur. J.
Wood Prod., 75, 807–815, 2017.
69. Ricci, A., Parpinello, G.P., Palma, A.S., Teslic, N., Brilli, C., Pizzi, A.,
Versari,  A., Analytical profiling of food-grade extracts from grape (Vitis
vinifera sp) seeds and skins, green tea (Camellia sinensis) leaves and
Limousin oak (Quercus robur) heartwood using MALDI-TOF-MS, ICP-MS
and spectrophotometric methods. J. Food Comp. Anal., 59, 95–104, 2017.
70. Ricci, A., Parpinello, G., Schwertner, A.P., Teslic, N., Brilli, C., Pizzi, A.,
Versari, A., Quality assessment of food grade plant extracts using MALDI-
TOF-MS, ICP-MS and spectrophotometric methods. J. Food Comp. Anal.,
59, 95–104, 2017.
71. Jahanshahi, S., Pizzi, A., Abdolkhani, A., Doosthoseini, K., Shakeri, A., Lagel,
M.C., Delmotte, L., MALDI-TOF and 13C-NMR and FT-MIR and strength
characterization of glycidyl ether tannin epoxy resins. Ind. Crops Prod., 83,
177–185, 2016.
72. Santiago-Medina, F.J., Foyer, G., Pizzi, A., Caillol, S., Delmotte, L., Lignin-
derived non-toxic aldehydes for ecofriendly tannin adhesives for wood pan-
els. Int. J. Adhes. Adhes., 70, 239–248, 2016.
73. Pizzi, A. and Roux, D.G., The chemistry and development of tannin-based
weather- and boil-proof cold-setting and fast-setting adhesives for wood.
J. Appl. Polym. Sci., 22, 1945–1954, 1978.
74. Pizzi, A., Rossouw, D.dU T., Knuffel, W., Singmin, M., “Honeymoon” pheno-
lic and tannin-based fast setting adhesive systems for exterior grade finger-
joints. Holzforsch. Holzverwert., 32, 6, 140–150, 1980.
75. Pizzi, A. and Cameron, F.A., Fast-set adhesives for glulam. For. Prod. J., 34, 9,
61, 1984.
76. Meikleham, N. and Pizzi, A., Acid and alkali-setting tannin-based rigid
foams. J. Appl. Polym. Sci., 53, 1547–1556, 1994.
77. Tondi, G., Pizzi, A., Olives, R., Natural tannin-based rigid foams as insula-
tion in wood construction. Maderas, 10, 3, 219–227, 2008.
78. Tondi, G. and Pizzi, A., Tannin based rigid foams: Characterisation and
modification, Ind. Crops Prod., 29, 356–363, 2009.
79. Tondi, G., Zhao, W., Pizzi, A., Fierro, V., Celzard, A., Tannin-based rigid
foams: A survey of chemical and physical properties. Bioresource Technol.,
100, 5162–5169, 2009.
80. Basso, M.C., Li, X., Giovando, S., Fierro, V., Pizzi, A., Celzard, A., Green,
formaldehyde-free, foams for thermal insulation. Adv. Mater. Lett., 2, 6, 378–
382, 2011.
81. Basso, M.C., Giovando, S., Pizzi, A., Celzard, A., Fierro, V., Tannin/furanic
foams without blowing agents and formaldehyde. Ind. Crops Prod., 49, 17–22,
2013.
82. Basso, M.C., Pizzi, A., Celzard, A., Influence of formulation on the dynamics
of preparation of tannin based foams. Ind. Crops Prod., 51, 396–400, 2013.
The Chemistry of Condensed Tannins 265

83. Basso, M.C., Pizzi, A., Celzard, A., Dynamic monitoring of tannin foams
preparation: Surfactant effects. Bioresources, 8, 4, 5807–5816, 2013.
84. Lacoste, C., Pizzi, A., Basso, M.C., Laborie, M.-P., Celzard, A., Pinus pinaster
tannin/furanic foams: Part 1: Formulation. Ind. Crops Prod., 52, 450–456,
2014.
85. Lacoste, C., Pizzi, A., Basso, M.C., Laborie, M.-P., Celzard, A., Pinus pinas-
ter tannin/furanic foams: Part 2: Physical properties. Ind. Crops Prod., 61,
531–536, 2014.
86. Basso, M.C., Pizzi, A., Lacoste, C., Delmotte, L., Al-Marzouki, F.A., Abdalle,
S., Celzard, A., Tannin–furanic–polyurethane foams for industrial continu-
ous plant lines. Polymers, 6, 2985–3004, 2014.
87. Basso, M.C., Giovando, S., Pizzi, A., Pasch, H., Pretorius, N., Delmotte, L.,
Flexible-elastic copolymerized polyurethane–tannin foams. J. Appl. Polym.
Sci., 131, 13, 2014.
88. Lacoste, C., Basso, M.C., Pizzi, A., Celzard, A., Ella Bang, E., Gallo, N.,
Charrier, B., Pine (P. pinaster) and quebracho (Schinopsis lorentzii/balansae)
tannin based foams as green acoustic absorbers. Ind. Crops Prod., 67, 70–73,
2015.
89. Li, X., Nicollin, A., Pizzi, A., Zhou, X., Sauget, A., Delmotte, L., Natural
tannin/furanic thermosetting moulding plastics. RSC Adv., 3, 17732–
17740, 2013.
90. Lagel, M.C., Zhang, J., Pizzi, A., Cutting and grinding wheels for angle grind-
ers with a bioresin matrix. Ind. Crops Prod., 67, 264–269, 2015.
91. Lagel, M.C., Hai, L., Pizzi, A., Basso, M.C., Delmotte, L., Abdalla, S., Zahed,
A., Al-Marzouki, F.M., Automotive brake pads made with a bioresin matrix.
Ind. Crops Prod., 85, 3, 372–381, 2016.
92. Pizzi, A., Tannin-based overlays for particleboard. Holzforsch. Holzverwert.,
31, 3, 59–61, 1979.
93. Abdullah, U.H., Pizzi, A., Zhou, X., Rode, K., Delmotte, L., Mansouri, H.R.,
Mimosa tannin resins for impregnated paper overlays. Eur. J. Wood Prod., 71,
153–162, 2013.
94. Abdullah, U.H., Zhou, X., Pizzi, A., Merlin, A., A note on the surface quality
of plywood overlaid with mimosa (Acacia mearnsii) tannin and melamine
urea formaldehyde impregnated paper: Effects of moisture content of
resin-impregnated papers before pressing on the physical properties of
overlaid panels. Int. Wood Prod. J., 4, 4, 253–256, 2013.
95. Abdullah, U.H., Pizzi, A., Zhou, X., High pressure paper laminates from
mimosa tannin resin. Int. Wood Prod. J., 5, 4, 224–227, 2014.
96. Pizzi, A., Kueny, R., Lecoanet, F., Masseteau, B., Carpentier, D., Krebs, A.,
Loiseau, F., Molina, S., Ragoubi, M., High resin content natural matrix-
natural fibre biocomposites. Ind. Crops Prod., 30, 235–240, 2009.
97. Sauget, A., Nicollin, A., Pizzi, A., Fabrication and mechanical analysis of
mimosa tannin and commercial flax fibers biocomposites. J. Adhes. Sci.
Technol., 27, 2204–2218, 2013.
266 Adhesives for Wood and Lignocellulosic Materials

98. Tondi, G., Wieland, S., Lemenager, N., Petutschnigg, A., Pizzi, A., Thevenon,
M.-F., Efficacy of tannin in fixing boron in wood: Fungal and termite resis-
tance. Bioresources, 7, 1, 1238–1252, 2012.
99. Tondi, G., Wieland, S., Wimmer, T., Thevenon, M.F., Pizzi, A., Petutschnigg,
A., Tannin-boron preservatives for wood buildings: Mechanical and fire
properties. Eur. J. Wood Prod., 70, 689–696, 2012.
100. Thebault, M., Pizzi, A., Dumarcay, S., Gerardin, P., Delmotte, L., Fredon, E.,
Polyurethanes from hydrolysable tannins obtained without using isocya-
nates. Ind. Crops Prod., 59, 329–336, 2014.
101. Thebault, M., Pizzi, A., Santiago-Medina, F.J., Al-Marzouki, F.M., Abdalla,
S., Isocyanate-free polyurethanes by coreaction of condensed tannins with
aminated tannins. J. Renew. Mat., 5, 1, 21–29, 2017.
102. Pizzi, A., Tannin: Major sources, properties and applications, Chapter 8,
in: Monomers, Polymers and Composites from Renewable Resources, M.N.
Belgacem and A. Gandini (Eds.), pp. 179–199, Elsevier, Amsterdam, 2008.
103. Krifa, M., El-Mekdad, H., Bentouati, N., Pizzi, A., Sick, E., Chekir-Ghedira,
L., Ronde, P., In vitro and in vivo anti-melanoma effects of Pituranthos tor-
tuosus essential oil via inhibition of FAK and Src activities. J. Cell. Biochem.,
117, 1167–175, 2016.
104. Krifa, M., El-Mekdad, H., Bentouati, N., Pizzi, A., Ghedira, K., Hammami,
M., El-Meshri, S.E., Chekir-Ghedira, L., Immunomodulatory and antican-
cer effects of Pituranthos tortuosus essential oil. Tumor Biol., 36, 5165–5170,
2015.
105. Krifa, M., Mustapha, N., Ghedira, Z., Ghedira, K., Pizzi, A., Chekir-Ghedira,
L., Limoniastrum guyonianum methanol extract induces immunomodula-
tory and anti-inflammatory effects by activating cellular anti-oxidant activ-
ity. Drug Chem. Toxicol., 38, 1, 84–91, 2014.
106. Noferi, M., Masson, E., Merlin, A., Pizzi, A., Deglise, X., Antioxidant char-
acteristics of hydrolysable and polyflavonoid tannins—An ESR kinetic study.
J. Appl. Polym. Sci., 63, 475–482, 1997.

You might also like