You are on page 1of 177

Prog. Polym. Vol.

22, 1203-1379, 1997


Sci.,
8 1997 ElsevierScience Ltd
Pergamon Printed in Great Britain. All rights reserved.
0079-6700/97 $32.00
PII: SOO79-6700(97)00004-X

FURANS IN POLYMER CHEMISTRY

ALESSANDRO GANDINI, MOHAMED NACEUR BELGACEM


Ecole Frayaise de Papeterie et des Industries Graphiques (INPG), BP 65, 38402
Saint Martin d’Ht?res, France

CONTENTS

1. Introduc don 1205


2. The furan heterocycle 1206
3. Sources of furanic monomers and reagents 1208
3.1. First-generation furans 1208
3.2. Furfural derivatives 1210
3.2.1. Furfuryl alcohol 1210
3.2.2. Furan, 2-methylfuran and tetrahydrofuran 1210
3.2.3. Furoic acids, furoates and difuranic coupled homologues 1211
3.2.4. 2-Furfurylamine and difurylic diamines 1212
3.25. Difuranic diisocyanates 1212
3.2.6. 2-Alkenylfurans 1213
3.2.7. 2-Furfurylidene methyl ketone and homologues 1213
3.2.8. 2-Furyloxirane and homologues 1214
3.2.9. 2-Vinyl furoate and homologues 1214
3.2.10. 2-Furfuyl acrylates and methacrylates 1215
3.2.11. 2-Fury1 vinyl ketones 1215
3.2.12. 2-Furfuryl vinyl ethers 1215
3.2.13. Furanic isocyanates 1215
3.3. Hydroxymethylfurfural derivatives 1216
3.3.1. 2,5Bis(hydroxymethyl)furan 1216
3.3.2. 2,5Bis(formyl)furan 1216
3.3.3. 2,5Furandicarboxylic acid, chloride and esters 1216
3.3.4. 2,5-Bis(aminomethyl)furan 1217
3.3.5. Furanic diisocyanates 1217
3.3.6. 2,5-Bis(chloromethyl)furan 1217
3.3.7. 2,5-Hydroxymethylfuroic acid 1218
3.3.8. 2-Hydroxymethyl-5-furanacrylic esters 1218
3.3.9. 2,5-Furandiacrylic acid 1218
3.4. Conclusions 1218
4. Step polymerisations 1219
4.1. The ’ ‘resinification” of furfural 1219
4.2. Furfural-based resins 1223
4.3. Furfuryl alcohol 1224
4.4. Furanic polyesters 1232

1203
1204 A. GANDINI and M. N. BELGACEM

4.5. Furanic polyamides 1238


4.6. Furanic polyurethanes 1246
4.7. Furanic polyschiff bases 1256
4.8. Furanic polyimidazoles 1261
4.9. Furanic polyethers 1263
4.10. Miscellaneous systems 1263
4.11. Conclusions 1264
5. Chain polymerisations 1265
5.1. Radical polymerisation 1266
51.1. The copolymerisation of furan and 2-methylfuran with maleic anhydride 1267
5.1.2. Copolymerisation of furans with acrylonitrile and other acrylic monomers 1268
5.1.3. 2-Alkenylfurans 1271
5.1.4. Furfuryl acrylate and methacrylate 1273
5.1.5. 2-Viny1 furoates 1274
5.1.6. Miscellaneous furanic monomers 1278
5.1.7. Retardating and inhibiting role of furans 1278
5.2. Cationic polymerisation 1280
5.2.1. Furan and alkyl derivatives 1280
5.2.2. 2-Alkenylfurans 1284
5.2.3. 2-Vinylfuroate 1297
5.2.4. Furfural and homologues 1297
5.2.5. 2-Furfuryl viny1 ether 1302
5.2.6. 2-Furfurylidene methyl ketone and homologues 1303
5.2.7. Miscellaneous furanic monomers 1306
5.3. Anionic polymerisation 1307
5.3.1. Furan and homologues 1307
5.3.2. 2-Alkenylfurans 1307
5.3.3. Furfural 1309
53.4. 2-Fury1 viny1 ketone 1309
5.3.5. Furanic isocyanates 1311
5.3.6. 2-Furfurylidene ketones 1312
5.3.7. 2-Furyloxirane 1315
5.4. Stereospecific polymerisation 1320
5.5. Conclusions on polyadditions 1321
5.6. Polymer modifications based on furan chemistry 1321
6. Conjugated furanic oligomers and polymers 1323
6.1. Poly(furan)s and copoly(furan)s 1325
6.2. Poly(2,5_furylene vinylene) 1327
6.3. Polyenes bearing lateral furan rings 1333
6.4. Conclusions 1335
7. Photopolymerisation and photocrosslinking 1336
7.1. Photopolymerisations from small molecules 1336
7.2. Photocrosslinking from chromophores on polymer molecules 1338
7.3. Conclusions 1346
8. Furan polymers and the Diels-Alder reaction 1346
8.1. Step polymerisations based on Diels-Alder cycloadditions 1349
8.2. The Diels-Alder reactions applied to furanic polymers 1354
8.2.1. Chemical modifications with monodienophiles 1354
8.2.2. Reactions with bisdienophiles 1357
8.3. The ROMP applied to furanic Diels-Alder adducts 1364
FURANS IN POLYMER CHEMISTRY 1205

8.4. Conclusions 1368


9. The ageing of furanic polymers 1369
10. General conclusions 1370
References 1371

1. INTRODUCTION

Twenty years have now passed since one of us published a review on this topic I and since
then a large number of studies related to it have appeared, justifying a comprehensive survey
of the field. Why specifically furans? Two reasons explain the logical restriction to such a
limited domain: (1) this heterocycle has a very peculiar chemical behaviour which is often not
closely related to that of its homologues thiophene and pyrrole, let alone aliphatic and
aromatic compounds; and (2) furan derivatives can be prepared from saccharidic renewable
resources and are therefore unique in terms of ubiquitous availability (there is biomass
virtually everywhere in the world) and continuous replenishing. The combination of these
unrelated aspects places furan compounds and polymers in a special context which motivates
organic, pharmaceutical and polymer chemists, materials scientists and process engineers.
Most of the research and development related to polymer science and technology in the
decades following World War II concentrated on petroleum-derived monomers. The use of
fossil raw materials was not considered problematic because of the huge reserves of these
resources. Today, the various hypotheses concerning their progressive depletion place the
onset of scarcity (at least for energy needs) at between two and four generations. It seems
therefore surprising that the alternative strategy, based on the exploitation of renewable
resources, namely the biomass, has not yet received the attention it deserves. Of course,
much is being done in the context of the production of energy from these natural products,
but there is still little concerted effort devoted to their use as building blocks for polymeric
materials. In other words, although some laboratories are engaged in this type of research,
mostly on their own initiative, what is lacking is a will to organise a long-term programme at
national or international levels. Furan chemistry is just an example of the application of this
working hypothesis which finds one specific outlet in the synthesis of polymers based on
these heterolcyclic structures, l-4 i.e. a rational transfer of solar energy into materials without
extinguishing the irreplaceable fossil inheritance. Other ways,2*3 falling outside the scope of
this review, include the chemical modification of cellulosics and other polysaccharides, the
use of lignins and vegetable oils as macromonomers or additives and the polymerisation of
various monomers obtained from the biomass, directly (e.g. terpenes) or through simple
chemical transformations (e.g. polyols from sugars).
The present survey concerns macromolecular structures bearing furan rings (or moieties
arising from them) because (1) furanic monomers or comonomers were used in their synth-
esis, or (2) furan derivatives were used as initiating, transfer or terminating agents, or (3)
conventional polymers were modified by furanic reagents. Each type of operation is relevant
here since the purpose of this review is to show not only the merits of furanic polymers per se,
but also the importance of exploiting the specificities of furan chemistry to elaborate original
structures and thus obtain interesting properties, even if this involves a very modest quanti-
tative intervention or presence of furan rings.
1206 A. GANDINI and M. N. BELGACEM

The literature quoted covers publications and patents available up to the end of 1996: a
deliberate and inevitably subjective selection was made in the choice of papers referenced
within the vast amount of material available, to the detriment of what was judged irrelevant to
the purpose of the review or too shaky in terms of scientific standing. Moreover, although
many publications already quoted in the previous review ’ will be re-analysed here in the light
of present knowledge, because of their permanent validity, much of what is now considered to
be scientifically obsolete, was left out. The historical value of these aged investigations will
be kept since they are discussed in the original survey. ’
The scope of the present review is strictly limited to polymer science and does not cover
other very important aspects of furan chemistry related to fine chemicals, synthons, etc.5-10
Suffice it to mention here one example of an extremely important application of a furanic
derivative, namely ‘ ‘ranitidine’ ’ , which is a very efficient hystamine type-2 receptor antago-
nist used as an anti-ulcer drug. This medicine is produced at a scale of about 500 tonnes per
year.

Ranitidine

2. THE FURAN HETEROCYCLE

Furan (1) is a well-known representative of a series of five-membered unsaturatred


“heteroaromatic” compounds which include many others, like pyrrole (2) and thiophene
(3). Each of these members is moreover the parent of a vast family of derivatives, and furanic
compounds are those relevant to this review.

4 3 <P>
510 \2(a)
X

1.x=o,furan
2. X = NH, pyrrole
3. X = S, thiophene

According to resonance theory and molecular orbital calculations, 1 possesses four reso-
nance-contributing structures as follows:

cl
I \ -0 - 0 4---b / ,O -
0 0ii 0 +--+-ooQ 0
:
1 la lb lc
FURANS IN POLYMER CHEMISTRY 1207

In furan chemistry, substitution reactions such as alkylation, halogenation, sulfonation and


nitration occur regioselectively at C2 and/or C5 when these positions are not substituted,
which suggests that la is the dominant resonance structure.
Compared with its homologues 2 and 3, 1 displays the lowest aromatic and the highest
dienic character, as illustrated below:

Increasing dienic character

Dewar resonance energies in kJ mol-l

95 27 22 18

Increasing aromaticity

The molecular dimensions of 1, which is a planar molecule, have been determined5 from
electron diffraction data:

a
a = 1.46 8. a= 107+4O
b = 1.35 8, p = 109 f 3”
c = 1.41 I!I0.02 A y= 107f.2”

1 is a colourless liquid with an ethereal odour, soluble in most organic solvents (such as
alcohols, ether, acetone). Its main physical properties are:

boiling point: 32”C/760 mmHg


freezing point: - 85.6”C
refractive index: 1.4215
density: 0.93

The chemical behaviour of furan and its derivatives displays a variety of features which are
associated with both the structure and properties of the heterocycle and the specific nature of
the substitue:nts appended to it. Classical5 and more recent6 comprehensive treatments of this
subject are available. The aspects relevant to this review will be emphasised in the corre-
sponding sections. Suffice it to say here that the mixed aromatic-dienic character comes to
light very frequently in the polymerisation mechanisms and that the predominance of the
latter feature makes many furans ideal substrates for the Diels-Alder reaction with a variety
of dienophilles. Conversely, the low extent of aromaticity often conspires against the stability
of certain fmanic derivatives, which tend to lose their original structure either through ring
opening or by giving rise to the corresponding dihydro- or tetrahydrofuran rings.
1208 A. GANDINI and M. N. BELGACEM

3. SOURCES OF FURANIC MONOMERS AND REAGENTS

This section is devoted to a brief survey of the furanic derivatives which are involved in the
polymerisation or polymer modification procedures that will be discussed later. No details are
given of their synthesis or properties, which are to be found in the specific references
provided in the appropriate sections or in more general reviews on furan chemistry and
compounds. sS6

3.1. First-generation furans

Before tackling the elaboration of macromolecules, it is important to deal with the synth-
esis of the small furanic molecules that will serve as building blocks or as promoters of
specific reactions. Most of these compounds are prepared from two basic structures which
constitute the first generation and which are obtained from saccharidic renewable resources,
namely 2-furancarboxaldehyde (commonly called furfural, 4) and 5-hydroxymethyl-2-fur-
ancarboxaldehyde (5).

HO4H

0 0
(4) (5)

Furfural is an “old” molecule since it was first obtained early last century5 and has
interested chemists ever since. It has been an industrial commodity for many decades because
it can be prepared quite readily and economically from a vast array of agricultural or forestry
wastes, all containing pentoses in sufficient amounts to justify a commercial exploitation:
corn cobs, oat and rice hulls, sugar-cane bagasse, cotton seeds, olive husks, wood chips, etc.
This peculiar state of affairs explains why the world production of 4, which is around
200000 tonnes per year, reflects in fact a spread of industrial activity (albeit in relatively
modest productions) involving many countries, including some which otherwise have little
infrastructure. Yields of up to about 15% of 4 with respect to dry matter are obtained from
processes which always involve an aqueous acidic medium and fairly high temperatures.“*
The reaction sequence goes first through the hydrolysis of the polymeric pentoses down to the
corresponding monosaccharide (aldopentose) which is then dehydrated progressively and
finally cyclised to give 4 (from xylose) and, in much smaller proportions, 5methylfurfural
(6) (from rhamnose):

OH OH OH OH
O+&OH -“+I*, HO+OH 420.

OH OH R OH OH R
FURANS IN POLYMER CHEMISTRY 1209

OH OH
o+JyH -%?
H R

0
(4) R = H
(6) R = CH3

As for 5, the interest in this molecule has been somewhat hampered by the difficulties
associated with its synthesis and storage, since it has a pronounced tendency to resinify as it
forms or on standing. Its preparation follows the same principles sketched above for 4, except
that the sources are now hexoses in the form of polysaccharides, oligosaccharides or simple
sugars. ‘-lo The classical precursor is fructose which is converted into 5 by acid-catalysed
dehydration--cyclisation as follows:

OH OH

HO+ 0
_3 HO
OH OH
OH OH
f i
HO HO HO

I -H20,H30+

0
H
H
H30+ 0 H,O+
H,O+
\ -g+
-H20
OH

-i
HO HO
1210 A. GANDINI and M. N. BELGACEM

- HOJ-&H
0
(5)

5 is not an industrial commodity at present, but pilot plants have been operating for the last
several years according to new processes which give high yields and the possibility of safe
storage. It is therefore reasonable to consider 5 as a future commercial source of fine che-
micals and monomers.

3.2. Furfural derivatives

3.2.1. Furjkyl alcohol

Most of the furfural produced worldwide is converted into furfuryl alcohol (7) by simple
reduction processes.’ In fact, at present, this is the most useful furanic compound, being used
to produce various types of resin as discussed in detail below. 5-Methylfurfuryl alcohol (8) is
synthesised likewise from 6.

&OH

Furfuryl alcohol can be converted into 2,5-bis(hydroxymethyl)furan (9) by formylation.’

OH + CH,O -H,O+
HO&OH

(9)

3.2.2. Furan, 2-methylfuran and tetrahydrofiran

The decarbonylation of furanic aldehydes 4 and 6 provides a source of 1 and 2-methylfuran


(lo),
respectively.’ Reduction of the former yields tetrahydrofuran (ll), which is of course
widely used as a solvent, but also as a monomer. This way of arriving at 11 represents an
alternative route to the petroleum-based process involving the dehydration of 1,4-butanediol.

/\
0, 0
0 0
(10) (11)
FURANS IN POLYMER CHEMISTRY 1211

The cationic polymerisation of 11 is outside the scope of the present survey because the
structure of this monomer has lost all the features characteristic of the furan heterocycle.
Excellent reviews on the subject are available1’V12covering both fundamental and applied
aspects and describing the means to prepare poly(tetramethylene oxide) glycols with a wide
range of degree of polymerisation (DP), which are important precursors for the preparation of
polyurethane elastomers.
The high susceptibility of the furan ring to electrophilic substitution at C2/C5 has been
extensively Iexploited. One classical application 13*14
is the condensation reaction of 1 or 10
with aldehydes and ketones in acidic media which affords a series of interesting compounds
ranging from di-2-furylmethane (12) to linear and cyclic oligomers, like 13 and 14, bearing
several heterocycles. The application of these concepts to polymer synthesis will be dealt
with in later sections.

R, RI and % = H, alkyl etc.

R, RI = H, alkyl etc.

3.2.3. Furoic acids, furoates and difuranic coupled homologues


The oxidation of 4 gives the corresponding 2-furoic acid 15, which can in turn be readily
esterified. These esters 16 can be coupled by a procedure15 similar to the condensation
reaction discussed above to give the difuranic diesters 17:
1212 A. GANDINI and M. N. BELGACEM

RO

R,, R2 = H, alkyl (17)

3.2.4. 2-Fwfwylamine and difuranic diamines

The conversion of 4 into 2-furfurylamine (18) is a conventional operation. What is relevant


to the present review is recent research devoted to the coupling of the corresponding chloro-
hydrate with aldehydes ” and ketones I7 to give the difmylic diamines 19:

CL
” 0
NH,
R2 R,

(la) (19)
a. R, = R, = H
b.R,H,=q=CH3
c. R, =R,=CH,
d.R,=H,R,=C,H,

3.2.5. Difuranic diisocyanates

The diamines 19 can be converted into the corresponding diisocyanates (20) using phos-
gene l8 or, better, triphosgene I9 (bistrichloromethyl carbonate):
FURANS IN POLYMER CHEMISTRY 1213

(20)
a. R, = R, = H
b.R,H,=R,=CH,
c. R,=R,=CH,
d. R, = H, R2 = C,H,

3.2.6. 2-Al$enyljiirans
2-Vinylfuran (21a) and its methylated homologues (21b-d) have been synthesised accord-
ing to different routes*“*’ and the characterisation of their structures and physical properties
is well documented. 20*22

R
(21)
a.R=R’=H
b.R=H,R’=CH,
c. R=CH,,R’=H
d.R=R’=CH,

3.2.7. 2-Furfurylidene methyl ketone and homologues

These compounds (22) are readily obtained by the base-catalysed condensation of furan-
carbonyl derivatives with aldehydes and ketones in aqueous media according to the well-
known aldol-crotonic reaction:

+ R -H,o- R’kR
Y
0 0 0
(22)
a.R’=H,R’=CH3
b.R’=H,R=CF,
d. R’ = H, R = t-Bu
e.R’=H,R=Ph
f.R’=R=H
g.R’=R=CH,
1214 A. GANDINI and M. N. BELGACEM

With an excess of the furanic reagent, the corresponding difurfurylidene derivatives, 22h and
22i, can be prepared in good yields, The synthesis of 2-furfurylidene methyl ketone (22a) and
its difuranic homologue (22h) has been studied in great detail,23 whereas other members of
the group have received more sporadic attention in the realm of polymer chemistry.24

h.R’=H
i.R’=CH3

3.2.8. 2-Furyloxirane and homologues

Compounds bearing furan and oxirane rings directly attached to each other are a relatively
recent addition to the family of furanic monomers. 2-Furyloxirane (23a) can be prepared from
4 in high yields by a simple one-pot procedure25 which has aIso been applied to 6 and to 2-
fury1 methyl ketone: x

R + (C-j,)++ I- NaoH’ H20. R’


- (CH,)+
0 0
(23)
a. R=R’=H
b. R = H, R’ = CH,
c. R=R’=CH3

3.2.9. 2-Vinylfuroate and homologues

Transvinylation between 2-furoic acid or its Ssubstituted homologues and vinyl acet-
ate 27*28provides a good entry to 2-vinyl furoates (24):

+ Od Ok4
0
‘r( 0
0

(24)
a.R=H
b.R=CH3
c. R = t-Bu
FURANS IN POLYMER CHEMISTRY 1215

3.2.10. 2-F~ji~yl acrylates and methacrylates


Acrylic and methacrylic esters of the furan series (25) are prepared by standard methods
involving the relevant acid chloride and alcohols:

(25)
a.R=H
b. R=CH,

These monomers have been the subject of some of the earliest studies on furan polymers.

3.2.11. 2-F1ql vinyl ketones


These compounds, e.g. 26, can be synthesised from the corresponding fury1 alkyl ketones:29

OK‘
I \
0
0
\

(26)

3.2.12. 2-F1~jiql vinyl ethers

The transvinylation reaction applied to 7 or 8..,.


and an alkyl vinyl ether gives the corre-
sponding furfuryl vinyl ethers 27 in good yields:‘”

(27)
a.R=H
b.R=CH3

3.2.13. Furanic isocyanates


2-Fury1 isocyanates (28) are very reactive compounds which must be stored at low tem-
peratures and preferably in solution. 19,31They are best prepared from the corresponding
furoic acids 15 by the Curtius rcaction.31 2-Furfuryl isocyanates (29) are easier to handle
and can be prepared in good yields from the corresponding amines by using triphosgene. I9
1216 A. GANDINI and M. N. BELGACEM

N, heatb fl
-N, R 0 N=C=O
0
(28)
a.R=H
b.R=CH3
c. R= Br

-+ COCCI, or
.&N=GO
Cl,CO-$--OCCl,
0
(29)
a.R=H
b.R=CH3

3.3. Hydroxymethylfirfural derivatives


This section is devoted to furanic monomers or reagents bearing two functional groups on
one heterocycle, i.e. essentially structures that are suited for polycondensation reactions.
Obviously 5 is the ideal precursor to this type of compound. lo Note that some difunctional
difuranic monomers have already been described in the previous section.

3.3.1. 2,5-Bis(hydroxymethyl)furan

An alternative way of synthesising 9 consists in the reduction of 5. With respect to the


formylation of 7, this reaction is simpler, but industrially it would require the availability of 5.

3.3.2. 2,5Bis(formyl)furan

The partial oxidation of the primary alcohol function of 51° gives the corresponding
dialdehyde 30. Contrary to 5, this compound is a stable furan derivative that can be stored
indefinitely in its crystalline state.

3.3.3. 2,SFurandicarboxylic acid, chloride and esters

Total oxidation of both functional groups in 5” provides the corresponding diacid (31a)
FURANS IN POLYMER CHEMISTRY 1217

which can be converted to its acid chloride (31b) and corresponding esters (31~) by standard
techniques.

H&&H q-&R
0 0 0 0
(30) (31)
a.R=OH
b.R=Cl
c. R = O-alkyl

3.3.4. 2,5-Bis(aminomethyl)jiuan

This diamine (32) is not easy to prepare because both the corresponding aldoxime and
nitrile resist reduction, and therefore the yields are 10w.~’ Surprisingly, contrary to lb
32 tends to resinify when stored in bulk at room temperature and it is therefore advisable
to convert it to its chlorohydrate salt which is more stable.

(32)

3.3.5. Furanic diisocyanates

The Curtius reaction33 of the dinitrile prepared from diacid 31a gives the corresponding
diisocyanate, 33. Like its monofunctional homologue 28, but to a lesser extent, this com-
pound is susceptible to react rapidly with external agents and must therefore be stored in an
inert atmosphere in a dark and cold place. The furfurylic counterpart 34, prepared from the
diamine 32 by using triphosgene,19 is more easily handled because of its relative kinetic
stability.

O=,N&N=C=O
N=C=O

3.3.6. 2,5-Bis(chloromethyl)furan

The furanic diol9 can be converted into its chloride 35 by standard procedures. Contrary to
2-fur-fury1 chloride, which is notoriously difficult to store as an undiluted liquid because of its
tendency to iself-condense, the fact of having both C2 and C5 positions substituted confers to
35 a much better ability to be kept as a crystalline powder.
1218 A. GANDINI and M. N. BELGACEM

3.3.7. 2,5-Hydroxymethylfuric acid

The selective oxidation of the aldehyde function of 5 gives the corresponding hydroxyacid
36. 34

Ho&OH
a&c1 0

3.3.8. 2-Hydroxymethyl-5-furanacrylic esters


A Wittig-type reaction applied to the aldehyde function of 5 gives the S-furanacrylic esters
37 bearing a primary alcohol group at Ck3’

R = Alkyl
(37)

3.3.9. 2,5-Furandiacrylic acid

This compound (38) is readily synthesised from dialdehyde 30 through the classical
malonic-type condensation, typically applied to convert 4 into 2-furanacrylic acid.

Rot 0
(38)

3.4. Conclusions

The variety of furanic monomers available, following published methods, is quite astound-
ing. Moreover, many of the preparations reported in recent studies show quite viable
approaches in terms of high yields and straightforward procedures from the two first-
generation furanic milestones. This state of affairs confirms that a door is now open into
the field of furan-containing polymers. The systems described in the following sections
will point to the fact that, although a good deal of research has been and is being done,
the area still awaits wider investigations both in the fundamental and the applied sectors.
As for novel monomers, or better synthetic methodologies for existing ones, the need
will become more apparent when the studies on furanic polymers and their applications
will demand fresh sources and/or improvements in their synthetic pathways.
FURANS IN POLYMER CHEMISTRY 1219

4. STEP POLYMERISATIONS
This section deals with systems involving polycondensation reactions and with the struc-
ture and properties of the ensuing materials. Some of these systems represent clear-cut step
polymerisations with well-understood mechanisms and macromolecular architectures,
whereas others, often termed ‘ ‘resinifications’ ’ , are characterised by ill-defined reaction
pathways and consequently by polymer structures which are far from being straightforward.
Linear, branched and crosslinked topologies will be included in this survey, and each specific
category will be examined and, whenever possible, explained.

4.1. The “resinijicatidn” of furjbal


Furfural is particularly sensitive to resinification, a feature which has been known for
decades, 56 and the classical product of this generic process, induced by acids and bases
(including zeolites37) but also, to a lesser extent, by high temperatures in neutral conditions,
is a black insoluble solid. When aqueous acidic media are used, the condensation reactions
are accompanied by hydrolytic ring opening, which is a general feature of the furan hetero-
cycle. The latter reaction is best described as the conversion of the unsaturated cyclic struc-
ture into aliphatic open-chain products bearing functions which depend on the specific furan
derivative used. Thus, for example, with 2,5disubstituted furans, but also with the corre-
sponding monosubstituted homologues and with furan itself, the accepted sequence of events
is:

,/o\ I \
0 R
H30+
slow
*
R

/ + I-40 F + H,O+
Rfi 0” R fast

OH

RJ+_R
0
JYf 0
R

When initially anhydrous conditions are applied, 4 gives self-condensation products which
were characterised in systems involving a large sample of highly purified monomer, kept in
1220 A. GANDINI and M. N. BELGACEM

an evacuated and sealed flask and submitted to a prolonged thermal treatment at 10%250°C
in the dark.38 Two reactive intermediates were isolated and identified despite their low
concentrations, namely the difuryl ketonic aldehyde 39 and the trifurylic dialdehyde 40.
The overall mechanisms of their formation were, in both instances, attributed to the inter-
vention of three molecules of substrate, according to the following schemes:

3Qy -q&H +RoH


0
(39)

(40)

Compound 40 was undoubtedly responsible for further condensation reactions as confirmed


by the formation of both soluble oligomers and insoluble resins. The net result of this process
was a slow, but progressive, accumulation of the black crosslinked product, whereas the
concentration of all soluble intermediates did not vary after reaching steady-state conditions.
This dynamic equilibration between the formation of soluble coloured condensation products
and their subsequent precipitation induced by further growth, is well illustrated by the con-
stant value of optical density, attained after the initial accumulation of the intermediate
structures. Fig. 1 shows an example of this behaviour. The oligomers could be separated into
two fractions, respectively soluble in methanol and methylene chloride, which showed a
marked difference in the degree of conjugation. No evidence for ring-opening reactions
was detected and elemental analysis of the oligomers and final products indicated the loss
of water to the approximate extent of one mole per three and two moles of reacted 4,
respectively, as confirmed by the actual amount of water measured in the residual liquid.
These “resins” exhibited intense EPR signals with g values around 2.007, characteristic of
highly delocalised unpaired electrons.
All this evidence suggests a series of condensation reactions as those just depicted, but
moving towards polyfurylic structures in which tertiary hydrogen atoms become particularly
mobile. They can therefore leave these branched oligomers and generate radicals which are
strongly stabilised by the extent of conjugation, as shown in the hypothetical macromolecular
fragment 41:
FURANS IN POLYMER CHEMISTRY 1221

(41)
It is most likely that this general set of events also occurs in the acid- and base-catalysed
resinification of 4. In fact, a mobile hydrogen atom, such as the tertiary H attached to a carbon
bearing three fury1 moieties, can not only be abstracted in the form of a free radical (II), but
also as H- (acidic media) or H+ (basic media). Interestingly, 5 is hardly sensitive to this type
of resinification, a fact that confirms the major role of the H5 atom in the condensation
mechanisms, as opposed to H3 and H4.
Recent work on the products of the pyrolysis of furfural resins obtained by acid
catalysis39*‘3 confirms the general features proposed above. These authors also studied the
morphology and the adsorption characteristics of these materials.41
A different behaviour was observed when liquid 4 was submitted to in vacuu photolyses.42
Excitation with wavelengths ranging from 265 to 360 nm produced the same features, albeit

q 0 D430

D 2

1000

Time (h)

Fig. 1. Colour formation at different wavelengths for the self-condensation reaction of 4 at


100°C in vucuo and in the dark.
1222 A. GANDINI and M. N. BELGACEM

at different rates. The major product was a soluble orange resin, always accompanied by
minor amounts of furoin (42). The structural analysis of the former suggested that the
photochemical events, following the excitation of 4, consisted first of two alternative dimer-
isation reactions giving 42 and 43, followed by an oligomerisation reaction induced by the
“hot” (vibrationally excited) 43. This led to successive molecular additions at the C5 posi-
tion of the ground-state substrate, giving structure 44 with DP of about 5:

0
b
HO HO. 0

(44)

This resinous product crosslinked slowly when exposed to air. This feature is not surprising in
the light of the extreme mobility of hydrogen atoms linked to a difurylmethane structure (see
Section 4.3 below) bearing here additionally an OH group. Indeed, difuryl carbinols are so
reactive that they cannot be stored in bulk.
The presence of 42 among the products indicated that head-to-head coupling could also
take place at any step in the oligomerisation:

J&H++LJ+Q-&
o*
FURANS IN POLYMER CHEMISTRY 1223

This reaction is equivalent to a termination step and its occurrence explains the low DP values
obtained.
No chain reactions involving free radicals were detected in these systems and the above
photo-oligomerisation was not inhibited by the addition of NO.42 This confirmed the mole-
cular nature of the process which involved, at each initiation step, one excited furancarbonyl
site and a ,ground-state molecule in which propagation involved a vibrationally excited
oligomer attacking a substrate molecule.

4.2. Furfural-based resins

Whereas the resins obtained from 4 alone seem to find applications that are limited to their
carbonaceous structures, i.e. as adsorption powders, much work has been devoted to the study
of resins in which furfural is coupled with other reagents.4 The most utilised comonomers are
phenols, bisphenols and acetone. Typically, an oligomer or “prepolymer” is synthesised for
commercial use and its application involves in situ crosslinking during processing,4 just like
with more widespread resins such as formaldehyde-based compositions. The detailed
mechanisms of these polycondensations are, however, much less well established with 4,
except perhaps for the systems involving acetone. We already mentioned the one-to-one
condensation product 22a which is a monomer in its own right (see Section 3.2.7 below).
However, in a more technological context, when 4 is made to react with acetone in aqueous
alkali,23 two major products are in fact formed, namely 22a and the difuryl homologue 22h
resulting from a double condensation, and it is this mixture that is used to obtain black
insoluble resins through acid catalysis. These resins have found several applications (at least
in East European countries) as adhesives, and corrosion-resistant coatings and floors. The
major mechanism in these cationic polymerisations involves the unsaturations in 22a and 22h
external to the furan rings, but condensation reactions involving carbonyl groups and elec-
trophilic substitution at C5 also occur, albeit to a much smaller extent.43*44 Thus, for 22a, the
polymer structure is essentially given by 45 in the first phase of the reaction and only later
does crosslinking take place by the side reactions just mentioned. These findings, strongly
supported by spectroscopic evidence and composition data, are in contradiction with earlier
qualitative studies which favoured, without clear proof, the predominance of condensation
reactions.

-
l2Yl
-0

/O
n
(45)

Recent additions to the catalogue of furfural-based resins include compositions with co-
monomers like quinacetophenone coupled with various substituted benzoic acids,45 hydro-
xyquinoline46 and cardanol,47 which were examined in terms of their chelating properties
towards metal ions. As already pointed out,4 despite the large number of publications on
1224 A. GANDINI and M. N. BELGACEM

resins containing furfural, much remains to be done in terms of the scientific understanding of
the mechanisms of polymerisation. In fact, if their industrial success is lagging, it is not for
lack of quality related to the polymers themselves, but rather owing to the insufficient
characterisation of these materials.

4.3. Furficryl alcohol

The literature on the resinification of 7 and on the properties and applications of the
ensuing materials is quite impressive in terms of the sheer number of articles and
patents. ‘V2,4,7*48,49
Areas in which these polymers find a successful and sometimes irreplace-
able usage include metal-casting cores and moulds, corrosion-resistant coatings, polymer
concretes, wood adhesives and binders, sand consolidation and well plugging, materials
possessing low flammability and low smoke release, and carbonaceous products comprising
industrial graphitic electrodes. Although the major component of all these resins is 7, many
“comonomers” like phenols, formaldehyde, urea, 4, and 9 have been used to achieve or
improve specific properties. The polymerisation process requires acid catalysis and always
leads ultimately to a black crosslinked product. Many studies have been devoted to the
understanding of the set of mechanisms which are responsible for the formation of these
networks. The present discussion will be limited to the “simplest” systems, i.e. those invol-
ving 7 alone in an acidic medium.50
When 7 is submitted to acidic conditions, irrespective of the nature and strength of the acid,
the presence or type of solvent and the temperature of the reaction, it goes through a global
polymerisation process involving a series of major transformations which sooner or later,
depending on the specific experimental conditions, yield the black crosslinked material
mentioned above. There is no doubt that the first phase of this complex process consists of
condensation reactions between the OH group on the one hand and either an H5 atom of a
furan ring (head-to-tail structure, 46) or another OH group (head-to-head structure, 47), on
the other:

(46)
- C&O
t

( n = 0, 1,2, etc.)

This series of acid-catalysed condensation steps produces linear oligomers which therefore
can contain both types of bridges, namely methylene (46) and dimethyleneoxide (47) moi-
eties. It has been established that, under strongly acidic conditions, the latter structure has a
tendency to lose formaldehyde and revert to the former, as indicated in the simplified scheme
above. This phenomenon, coupled with the higher kinetic propensity for head-to-tail con-
densation, results in macromolecular structures in which methylene bridges are in general
much more frequent than their ether counterparts.
FURANS IN POLYMER CHEMISTRY 1225

It is quite obvious that if these two reactions were the only major events in the polycon-
densation of 7, the corresponding products would be colourless thermoplastic materials. In
fact, they would possess linear structures deprived of any chromophores (2,5_dialkylsubsti-
tuted furam absorb below 250 nm). However, such simple structures can only be isolated if
the reactions are stopped at the very early stages of the process, namely when average DP
values are between 2 and 5, as shown by the different compounds isolated and identified by
different gro~ps.~‘-~~ Even then, the actual mixture of products has a brown colour and
already contains more complex structures, including unexplained aliphatic moieties.53 When
the systems are allowed to evolve further, their behaviour changes drastically in that cross-
linking and darkening occur without the possibility of isolating intermediate branched pro-
ducts, since: no soluble material of medium or high DP has ever been characterised. In other
words, there is a sudden transition between mostly low-DP oligomers and black-coloured
networks. This feature has puzzled chemists for decades and only recent work has helped to
explain most of the mechanisms responsible for its occurrence.50
The qualitative difference between the numerous previous studies which attempted to
unravel this problem and our investigation consisted in calling upon a large number of model
compounds. which progressively excluded the likelihood of certain mechanisms proposed
earlier and pointed more and more clearly towards the real causes of the observed phenom-
enology. The first set of models involved various furanic alcohols in which the primary
character of 7 was progressively changed into secondary and tertiary. It was found that 48
behaved quite similarly to 7 in the acid-induced resinification, albeit at lower rates, whereas
49 displayed a very different behaviour giving linear oligomers (DP up to cu. 10) which were
practically colourless. The regular structures thus obtained are clearly inferred from Figs 2
and 3, which show the FTIR and the ‘H-NMR spectra of a typical oligomer of 49.

This radical change in reactivity suggests that hydrogen atoms attached to the carbon
linked to two furan rings must play a key role in the mechanisms leading to both crosslinking
and coloration. Previous reports had often claimed that the C3 and C4 positions of the
heterocycle were responsible for the former mechanism through condensation reactions
similar to Ithose encountered in the linear growth. This has now been ruled out,50 not only
because the products of the polycondensation of 49 are linear, but also because a series of
experiments with 8 only gave the dimer 52 resulting from head-to-head condensation and the
loss of formaldehyde.
The other aspect which has been definitively cleared up concerns the role of water and the
hydrolysis products that an acidic aqueous medium can produce from furanic derivatives, as
discussed in Section 4.1. It had been proposed that some of these aliphatic structures were the
cause of crosslinking. However, more recent experiments using thoroughly dried media with
2-furfuryl acetate (50) and with a mixture of 7 and its silylated derivative 51 showed5’ the
occurrence of all the general features observed with 7, which suggested that the absence of
1226 A. GANDINI and M. N. BELGACEM

-
4000 3000 2000 1500 1000 500

Wavenumber (cm-‘)

Fig. 2. FTIR spectrum of a typical oligomer of 49. (Reproduced by permission of the


American Chemical Society. Copyright 1996. Reprinted from “.)

1 6 5 4 3 2 I 0

pm

Fig. 3. ‘H-NMR spectrum of a typical oligomer of 49. (Reproduced by permission of the


American Chemical Society. Copyright 1996. Reprinted from ‘O.)
FURANS IN POLYMER CHEMISTRY 1227

water did not alter the process and in particular did not stop the formation of network
structures.

The second set of models was aimed at investigating the reactivity of the linear oligomers.
For this, the three dimeric structures 52-54 were prepared and submitted to different me-
chanistic tests. 5oThe external C5 positions were blocked by methyl groups to avoid electro-
philic chain-extension reactions.

(52)

260 320 380 440 500 560 260 320 380 440 500 560 260 320 380 440 500 560
Wavelength, nm Wavelength, nm Wavelength, nm

Fig. 4. UV-vis. spectra of (a) 52, (b) 53, and (c) 54: A, lo4 M solution in CH2C12 of the
corresponding model; B, A plus an equal volume of a 10” M solution of the hydride-ion
abstactor; C, B after neutralisation with triethylamine. (Reproduced by permission of the
American Chemical Society. Copyright 1996. Reprinted from 50.)
1228 A. GANDINI and M. N. BELGACEM

The reaction of these compounds with an excess of hydride-ion abstracters, like dioxolenium
or trityl ions, was followed by both electronic and ‘H-NMR spectroscopy. With 52 and 53 the
formation of the corresponding carbenium ions by loss of H- from the central carbon was
unambiguosly established, whereas 54 did not react with either abstractor and was recovered
unchanged. Fig. 4 shows examples of the UV-vis. spectra arising from these interactions. In
Fig. 4(a), model 52 was treated with a dioxolenium salt and the peak at 490 nm reflected the
formation of the corresponding difurylmethyl carbenium ion which is bleached by the
addition of triethylamine. Fig. 4(b) shows the same phenomenology applied to 53, which
can lose its only hydrogen atom in the same manner. Fig. 4(c) clearly indicates that 54
was inert towards hydride-ion abstraction. The ‘H-NMR spectra of these systems were
taken with much more concentrated solutions and confirmed the occurrence of the H-
transfer from 52 and 53 (formation of the dioxolane derivative or PhsCH) and the lack of
reactivity of 54.
These mechanistic tests corroborated the conclusion on the essential role of mobile hydro-
gen atoms in favouring the side reactions related to colour formation during the resinification
in an acidic medium. However, in order to explain the presence of a deep colour in the
neutralised polymers, it became necessary to prepare model compounds with higher DP
values. A trimer like 52 and linear oligomers obtained in polycondensations of 7, stopped
at an early stage, were used to show that indeed a sequence of events leading to highly
conjugated structure 55 took place by successive hydride-ion/proton abstractions, as shown
in the following general scheme: 5o

[Y _ : (SnCI,OH) _ ou OH,A’ ]

1 -HY
FURANS IN POLYMER CHEMISTRY 1229

or
+
_~__________________~~~~~~~~~

+ (Hj ‘1

(55)

Interestingly, this overall mechanism also explains the presence of methyl groups observed
by several authors in the oligomers of 7.51-53The reaction of these oligomers with hydride-
ion abstracters gave UV-vis. spectra that clearly confirmed the existence of different con-
jugated sequences, because the ensuing carbenium ions absorbed at various long wave-
lengths, as shown in Fig. 5.
All the above experiments and ensuing mechanistic proposals are limited to the appearance
of chromophores like 55, but do not explain the crosslinking of the oligomers. Having ruled
out the possible intervention of the C3 and C4 positions, as well as of the hydrolysis products,
it follows inevitably that the non-furanic unsaturations arising from the H-/I-I+ loss cycles are
the only possible source of branching. Given the very sudden character of the formation of
insoluble products, it seems moreover logical to assume that the crosslinking involves inter-
molecular events among macromolecules. One way to rationalise these requirements is to
postulate Diels-Alder cycloaddition reactions between a furan moiety (diene) and an exo-
unsaturated dihydrofuranic cycle (dienophile),50 e.g.:

/ -O/ O
\ / _ ’ \O/
1230 A. GANDINI and M. N. BELGACEM

(56)
This mechanism is at odds with yet another previous proposal that contemplated a condensa-
tion reaction between a methylene group bridging two furan rings and an OH end group from
an oligomer or polymer chain:54

- H20 JJ&Q%
+

OH TiCl,

(57)

I
3’ 436 552 668 784 9
Wavelength (nm)

Fig. 5. UV-vis. spectra of an oligomer of 7: A, 50 mg 1-l solution in CH2ClI;B, A plus an


equal volume of 10” M solution of the hydride-ion abstactor; C, B after neutralisation
with triethylamine. (Reproduced by permission of the American Chemical Society. Copy-
right 1996. Reprinted from ‘O.)
FORANS IN POLYMER CHEMISTRY 1231

The argument against this proposal, which was based on solid-state NMR data, comes from
experiments in which we attempted to simulate this reaction with model compounds, i.e. 52,
plus 8 under typical resinification conditions. The products isolated were just 52 in amounts
larger than those introduced, and a brown soluble aliphatic resin.50 The only relevant reac-
tions occurring were therefore the head-to-head self-condensation of 8 followed by loss of
formuldehyde to give 52 and the acid-catalysed hydrolysis of the substrates by the water
produced k the first interaction. The hydrolysis products were then turned into the aliphatic
resin. No evidence of structures like 57, arising from the condensation of 52 with 8, was
obtained in these tests. The essential difference between the two mechanisms leading to chain
branching and ultimately crosslinking resides in the fact that, whereas the formation of tbe
Diels-Alder adcluct 56 requires the presence of previously formed dienophilic unsaturations,
Le. of conjugated sequences like 55, the condensation scheme only involves methylene and
hydroxy functions which are present at all times once the polymerisation has started. Thus, in
one case the branching is conditioned to the H-/I-I+ cycles having occurred, whereas, in tbe
other, the two events are independent. As alread emphasised, none of the studies devoted to
the isalation and characterisation of oligomers z-53 ever detected any non-linear structure,
which tends to invalidate further the simple -Fu-CHr-FU- + -Fu-CHzOH condensation
mechanism in favour of the one proposed here.”
Given the high degree of conjugation of the final products 56, it is not surprising that
various autbors have reported the presence of paramagnetic centres by ESR measurements.55
These should, however, not be attributed to a specific radical reaction during the polycon-
densation process, but rather to oxidation of the black crosslinked material after exposure to
air.
Having reached a satisfactory understanding of the causes of the two major side events in
the resinification of 7, it would of course be extremely interesting to find ways to minimise
them in order to prepare light-coloured thermoplastic materials from this monomer. To date,
no solution has been proposed to this challenge.
The soluble oligomers of 7 have been used not only as viscous precursors to the final
crosslinked resins employed in the applications mentioned above, but also as co-reagents for
the preparation of different macromolecular materials. Thus cellulose can be grafted by these
oligomers tbrough a classical free-radical processs6 although the specific mechanism of their
attachment to the polysaccharide backbone remains obscure.57 In another vein,58 acrylonitrile
was copolymerised with these oligomers by free-radical means and it was shown that their
incorporation (giving soluble products with up to 80% by weight of furanic structures)
occurred through the terminal furfuryl unit acting as a diene. A third recent investigation
deals with the formation of interpenetrating networks between elastomeric polyurethanes and
rigid crosslinked furfuryl alcohol structures.59*60The growth of the two networks was
chemically independent59 because the aminic agent provoking the crosslinking of the poly-
urethane was much more reactive towards NC0 groups than the OH functions borne by the
oligomers of 7. These materials were characterised thoroughly in terms of thermal and
mechanical properties, and tested both as such59 and as matrices for glass-fibre-reinforced
composites. 60
Another area of interest related to the use of 7 as a monomer concerns the coating of
inorganic particles and has been studied in particular with silica.6’ The solid phase is dis-
persed in ain organic solvent in which au acidic catalyst and 7 are added. The polycondensa-
tion takes place both homogeneously in the solvent and heterogeneously on the solid surface.
1232 A. GANDINI and M. N. BELGACEM

The structure of the polymer formed around the particles closely resembles that of the
standard crosslinked material discussed above.
Finally, much attention has been given recently to the pyrolysis of resins derived from 7
and to the various carbonaceous materials thus obtained like carbon molecular sieves,@
glassy carbon, 63 matrices for carbon/carbon composites,@ electrodes for both secondary
batteries65 and metallurgical processes, etc. The mechanism of pyrolysis has been looked
into by following the NMR spectra,66 the elemental composition and the IR spectra. 67Up to
about 300°C the condensation and crosslinking reactions dominated, whereas above about
35O”C,loss of hydrogen and oxygen led progressively to carbonisation which reached com-
pletion around 900°C.
If, on the one hand, the possibility of synthesising linear (CHz-2,5-Fu) type polymers from
7 is very attractive, but far from evident, the use of the well-known conjugated materials for
novel applications like electronic conduction is, on the other hand, much more readily
feasible and should therefore deserve closer inspection.

4.4. Furanic polyesters

Apart from sporadic studies describing the use of difunctional furanic monomers to prepare
polyesters, as reviewed earlier, 1*2*48
the first systematic approach to this topic was carried out
using both furanic diol 9 and diacid derivatives 31 (as well as 2,3-, 2,4- and 3,4-furanic
diacids) reacting with aliphatic or aromatic counterparts to give polymers like 58 and 59,
respectively, or between themselves to give the fully furanic structures 60. 68-72The ensuing
polyesters had relatively low DP values, as suggested by their low intrinsic viscosities. Only
one of the numerous structures studied, namely 58 with 2,5enchainments, exhibited a clear
melting feature, as observed by DSC. Thermal analyses also showed that these polyesters
differed in their degradation behaviour in that structures 59 were quite resistant to heat and
atmospheric oxidation, whereas structures 58 and 60 displayed a marked fragility under these
severe conditions. This difference arises, at least for 60, from the lability of the methylene
hydrogens connected to the furan ring, as opposed to the stability of the Fu-C=O moieties.
Another possible source of problems related to this work is the fact that the classical catalysts
and procedures adopted for the syntheses were probably not adapted to (some) furanic
monomers and consequently could have introduced some weaknesses (structural defects)
in the polyester chains.

w”-od 2,5-, 2,4- or 3,4-


substituted furanic ring
(58)
FURANS IN POLYMER CHEMISTRY 1233

_
2,5-, 2,4- or 3,4-
substituted furanic ring
(59)

2,5-, 2,4- or 3,4-


substituted furanic ring
(60)

The use of hydroxyacid 36 to prepare the corresponding furanic polyester only led to linear
(61a)
and cyclic (61b)
oligomers with DP values up to 5,when the monomer was treated with
pyridinium salts at 60°C.73 The proportion of the two products depended on the reaction
conditions, which were thoroughly scanned. Melt polycondensation of 36 or transesterifica-
tion with ethylene glycol gave black, intractable, resinous products. Despite the quality of this
study, this monomer deserves further attention within a wider scope with respect to the search
for more appropriate polycondensation catalysts, also applied to the corresponding hydro-
xyesters, as in the case of its homologue 37 discussed below.

n=3to6
(61a) (61b)

The hydroxyester 37 was found to be an interesting source of the corresponding furanic


polyester 62 via a transesterification reaction promoted by a mild catalyst, i.e. KHC03:74
1234 A. GANDINI and M. N. BELGACEM

(62)

The ethanol produced in the condensation was easily vacuum-removed and the reaction
temperature was raised progressively as the viscosity increased. This polymer did not grow
beyond DP values of about 20, because it crystallised out of the reaction medium both from
bulk and solution systems. Its structure was characterised thoroughly and was found to be
quite regular as shown in Fig. 6, which gives the ‘H-NMR spectrum of a typical product. The
crystalline form of polyester 62 had a melting point of 180°C and its glassy form had a Tg of
45°C. Fig. 7 gives the DSC traces related to these transitions. TGA analyses showed that this
polymer was stable up to 280°C. The presence of a strong chromophore in each monomer unit

Ho~oJ-&-#*yll
6
L
0 -In 0

I+8
3

r
11
6’ 10

8.b 7:s 7:o 5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0

Fig. 6. ‘H-NMR spectrum of polyester 62.


FURANS IN POLYMERCHEMISTRY 1235

(a>

130

(b) 45

\
180

1 w
0 50 100 150

Fig. ‘7.DSC traces of polyester 62: (a) first scan and (b) second trace after quenching.

along the chain provided a convenient source for photoactivation. The polymer, just like its
monomer, displayed in fact a maximum at 360 nm tailing off into the visible and giving
material of yellow colour. Upon irradiation with a high-pressure mercury lamp the polymer
crosslinked and turned white. A close inspection of the resulting network revealed that the
furan and carbonyl moieties had not been altered, but that the C% unsaturations had been
partly removed because of the occurrence of molecular photocycloaddition reactions invol-
ving an excited and a ground-state chromophore. This gave rise to the classical cyclobutane
unit:

The conjugation in these sites was therefore eliminated and the crosslinked polymer now
absorbed much less in the near UV. Moreover, the intensity of the IR peak of the C=C
unsaturation at 1640 cm-’ was drastically reduced as a result of this photocyclisation. This
decrease was accompanied by an increase in absorbtion in the aliphatic CH region around
1236 A. GANDINIand M. N. BELGACEM

100 _
I, “. I “. 1 ’ ‘. 1 ”
100 200 300 400

‘I- (“0

Fig. 8. TGA trace of polyester 63. (Reproduced by permission of Elsevier Science Ltd.
Copyright 1993. Reprinted from 75.)

2900 cm-’ and a shift of the C=O band towards higher frequencies because of the loss of
conjugation in the cyclised units.
This investigation74 represents the first report of both a crystalline and photosensitive
furanic polyester.
An interesting addition to the realm of polymers derived from biomass is a recent report75
on the synthesis of polyesters derived from the diacid dichloride 31b and the three isomers of
1,4:3,6-dianhydrohexitols (isosorbides), which are industrial products derived from certain
hexoses. The polycondensations were carried out at low temperature in 1,1,2,2-tetrachloro-
ethane using pyridine as a catalyst and a slight excess of 31b. The reactions proceeded
smoothly despite the difference in reactivity between the two OH groups in each isomeric
diol. The polyesters isolated by precipitation in methanol (e.g. 63) had number-average
molecular weights around 20000 and their elemental analyses and NMR spectra suggested
regular structures with a high macromolecular stiffness, as indicated by T, values around
180°C. Curiously, despite this apparent structural regularity, no mention of crystallinity or of
melting features by DSC or microscopy observation was made for these polyesters. Their
thermal stability was quite satisfactory since even in air no weight loss was detected below
32O”C, as shok in Fig. 8.

'0

%
0

OY?f
0

(63)
I \
0
0
0

n
FURANS IN POLYMER CHEMISTRY 1237

The use of difuranic diesters or of the corresponding acid chlorides to prepare polyesters
has been undertaken and for the moment only aliphatic diols have been employed as com-
plementary Imonomers,76to obtain e.g. structure 64. Both transesterification and direct con-
densation with a proton acceptor gave regular polymers with unoptimised molecular weights
of up to 7000 and T, values going from room temperature to 60°C and melting temperatures
around 7O”CIfor the higher molecular weight products. The interest in utilising these furanic
monomers is the possibility of varying their structure as a function of the precursor aldehyde
or ketone in their synthesis. This investigation is in progress.

m=2or4
(64)

Fully aromatic polyesters derived from terephthalic acid and substituted hydroquinones
have been studied extensively in view of exploiting their liquid-crystal properties, but in
general they have the drawback of excessive chain stiffness which translates into unattainable
melting temperatures, i.e. higher than 500°C. The replacement of the aromatic acid by
the 2,5-furanic counterpart 31 was studied briefly and led to products which, unfortu-
nately, were equally untractable because they were neither soluble nor meltable.77 The
use of the aromatic-furanic acid chloride 65 with 1,8-octanediol gave a polyester which had a
melting temperature of 19O”C,without clear indications of the presence of mesophases.77 In
order to verify whether the basic units related to furanic-aromatic polyesters had mesogenic
properties, model compounds bearing central (e.g. 66) or lateral (e.g. 67) furan rings were
synthesised and characterised by DSC.78 As suspected, the relatively low core angle asso-
ciated with the 2,5-furandicarbonyl structure, namely about 130”, was detrimental to the
formation of liquid-crystalline mesophases even when polar end groups were appended to
these structures.

(65)
1238 A. GANDINI and M. N. BELGACEM

n=2,3,4and6and7
(66)

&@-Q”*
0 0 0 0
(67)

These negative results do not necessarily imply that the corresponding polyesters will also
be deprived of mesophases, since the macromolecular character can induce a qualitative
change in that direction as with certain cellulose derivatives and with the furanic/aromatic
polyamides like 69 (see below).
What remains to be investigated are polymeric structures that are either soluble or have
more accessible melting temperatures. Some research has indeed been carried out in this
direction: because furan rings introduce distortions with respect to the rigid rod-like structure
of completely aromatic polyesters, copolyesters with up to 50% of 2,5-furancarbonyl moi-
eties were synthesised and characterised.79 It was found that the random presence of the
heterocyclic units produced a drop in the melting temperature of the product: thus in one
specific series, with a fully aromatic polyester, the T, was 415°C and the progressive
incorporation of furanic comonomer units gave a correspondingly lower T,, down to
340°C when equal amounts of phthalic and furanic acids were used.
Curiously, the thiophenic/hydroquinone combinations do lead to liquid-crystalline polye-
sters,” but in that case the core angle for 2,5-disubstituted thiophenes is around 150”. It seems
that the critical value of this angle, determining the feasibility of mesogenic structures, is
between 130” (2,5-furan) and 150” (2,5-thiophene). Indeed, polyesters based on hydroqui-
nones and 1,3-phthalic acid (core angle 120”) do not give a liquid-crystal behaviour.

4.5. Furanic polyamides


The synthesis of furanic polyamides has concerned polymer chemists since the early
1950s but reliable systematic studies with clear-cut structural determinations are few and
far between. ‘,2*48Most of this early work was devoted to the combination of diacid derivatives
31 and aliphatic or alicyclic diamines,8’-85 although other furanic diacids were also investi-
gated. s4*85The resulting polymers 68 were intrinsically soft, nylon-type materials because of
the flexibility of the oligomethylene moieties in the main chain.
FURANS IN POLYMER CHEMISTRY 1239

(68)
On the whole, these investigations did not pursue the details related to molecular weight,
crystallinity or physical properties and remained therefore rather qualitative in their
approach. Moreover, large discrepancies are observed from one publication to another in
some of the: data referring to the same expected polyamide structure, such as Ts, T,,, or
decomposition temperature, so that doubts remain about the reliability of some of these
studies. It can be concluded that, despite the relatively large number of publications,
furanic-aliphatic polyamides have not yet been studied in a comprehensive fashion.
Furanic polyamides built from 31 and aromatic amines were first reported in 1964, but
without any structural study.86 At a much later date, this type of material was re-examined in
a more detailed fashion. 87S88Both the acid 31a and its 3,4-homologue were used in combina-
tion with various aromatic diamines and the polymerisation procedures adopted included all
the classical methods, but also interfacial systems and an extension of the so-called direct
polycondensation technique developed previously for fully aromatic polyamides. The latter
type of synthesis gave the best results in terms of structural regularity and molecular weights.
Reactions were carried out under stoichiometric conditions in NMP/pyridine in the presence
of triphenyl phosphite, calcium and lithium chloride at 90°C for 20 h. The overall mechanism
leading to polyamides through “direct condensations” between COOH and NH2 groups is
given below in a schematic representation:

0
04
R, r’”
/’
A
-0Ph

H-i-OCOR, 0
Phd ‘OPh
K
R, OH

PhOH

Among the diamines used, the most promising was 1,6phenylene diamine because the
direct polycondensation requires the use of extremely pure monomers and that was the
compound most easily purified. Of course, the same considerations apply to the diacids
and in this case 31a was the monomer to which major efforts were devoted in order to
1240 A. GANDINI and M. N. BELGACEM

achieve a thorough purification. Thus, the “best” polyamide obtained in that study was
69 87988

3000

2000

z
.13
L
F

1000

C Wdl)

Fig. 9. Plot of the shear viscosity (measured at a shear rate of 2.84 s-l) as a function of
concentration of polyamide 69 in solution of NMP with 4% of LiCl, at 40°C. The sample
used had an inherent viscosity of 2.1 dl g-’ (measured in 98% sulfuric acid at 30°C with a
concentration of 0.5 g dl-‘). (Reproduced by permission of the American Chemical
Society. Copyright 199 1. Reprinted from s8.)
FURANS IN POLYMERCHEMISTRY 1241

(a) A

100 - -

B /A --
---- __,-----
40 -
I I 1 I 1 I
100 200 300 400 500 600
T V'C)

4
$

b
@I C

w
i

4
z I I I I 1 1

100 200 300 400 500 600


TV'0

Fig. 10. Thermal analysis of polyamide 69. (a) TGA curves at 5°C min-‘: A, weight loss;
B, rate of weight loss. (b) DSC trace at 10°C min-‘: curves C and D differ in sensitivity by
a factor of 2.5. (Reproduced by permission of Elsevier Science Ltd. Copyright 1985.
Reprinted from “.)

It was only soluble in concentrated sulfuric acid and NMP containing 4% LiCl, had repro-
ducible inherent viscosities of 2.1 to 2.3, and its ‘H- and 13C-NMR, as well as FI’IR spectra,
reflected a very regular structure. Molecular weights were high, i.e. M, close to lo5 and Mn
close to 20000, and with these high polymers the solutions displayed a lyotropic liquid-
crystal behaviour, as shown by shear viscosity measurements. Fig. 9 gives a typical plot of the
shear viscosity of one such solution. The dramatic drop in viscosity, corresponding to a
critical concentration at about 12%, indicated the occurrence of macromolecular alignments
associated ,with the formation of the lyotropic mesophase. DSC and optical microscopy
observations 88 confirmed this behaviour and, more quantitatively, led to the construction
of the isotropickmisotropic phase diagram relating temperature to polymer concentration.
1242 A.GANDINIand M.N.BELGACEM

L 1
v l
11 10 9 8 I 6 5 4
6 @pm)

Fig. 11. ‘H-NMR spectrum of polyamide 71a.(Reproduced by permission of the Amer-


ican Chemical Society. Copyright 1991. Reprinted from **.)

The thermal properties of polyamide 69, determined by DSC and TGAg7 as shown in Fig. 10,
displayed only a minor glass transition feature at 325°C below 39O”C, when the polymer
started decomposing. Thus, thermal degradation preceded melting with this structure, just as
with the corresponding fully aromatic polyamides like Kevlar.
Replacement of 31 by its 3,4-homologues7*** resulted in a less rigid chain as shown by an

Table 1. Thermal properties of furanic polyamides

Polyamide Onset of decomposition Tg W) T, (“C) Ref.


(“C)
69 385 325 87

70 380 205 - 87
71a 360 285 - 87

71b 315 - - 87
71c 365 - - 87

72 320 - - 87
73 325 - 325 88
74 260 100 180 88
FURANS IN POLYMER CHEMISTRY 1243

5 6

160 140 120 100 80 60

Fig. 12. 13C-NMR spectrum of polyamide 71b. (Reproduced by permission of the Amer-
ican Chemical Society. Copyright 1991. Reprinted from **.)

increased solubility and a T, of 205°C. This polyamide (70), whose synthesis was not
optimised and which had therefore lower molecular weights than 69, displayed however
essentially the same thermal stability as its 2,Sfuranic homologue.

Replacing 1,Cphenylene diamine in 69 with diaromatic counterparts87~88 provided a series of


furanic-aromatic polyamides 71 which again had regular structures as shown in the typical
spectra in Figs 11 and 12. Their solubilities and thermal properties (T,, T, and onset of
decomposition) varied with the nature of the spacer between the two benzene rings, as shown
in Table 1.
1244 A. GANDINI and M. N. BELGACEM

n
(7l)
a.R=S02,b.R=Oandc.R=CH,

Finally, the study of these polyamides was extended to include fully furanic structures by
using both 25 and 3,4-diacids and three furanic diamines.87s8 Because furanic amines in
which the amino group is attached directly to the heterocycle are thermodynamically unstable
and tautomerise to the corresponding imines, only monomers bearing two aminomethyl
moieties could be used in this investigation. This inevitably increases the flexibility of the
ensuing polyamides because the methylene groups now play the role of spacers at each
amidic site. The three polymers 72-74 were prepared in relatively low yields and gave
viscosities which suggested rather modest molecular weights, but their NMR and IR spectra
were entirely consistent with the expected structures. The poor quality of these syntheses
arose mostly because of the difficulty in obtaining the furanic diamines to the same level of

100 J
100 200 300

‘I- (“Cl

Fig. 13. DSC and ATG thermograms of fully furanic polyamide 72. (Reproduc;: by
permission of the American Chemical Society. Copyright 1991. Reprinted from .)
FURANS IN POLYMER CHEMISTRY 1245

60 140 220 300

T PC)

Fig. 14. DSC and ATG traces of fully furanic polyamide 74. (Reproduced by pit-mission of
the American Chemical Society. Copyright 1991. Reprinted from .)

purity achieved with the aromatic ones. Polymer 72 started melting at 3OO”C, but decomposed
soon after. Its 3,4-furanic homologue 73 gave a similar behaviour but around 240°C.
Polyamide 74 displayed a T, at 100°C and melted between 180 and 22X, with thermal
degradation starting at 260°C. Figs 13 and 14 show these features for polyamides 72 and 74,
respectively. These relatively low transitions and decomposition thresholds are related to the
flexibility iutroduced by the methylene and oxymethylene moieties, which are also a source
of lability as discussed in Section 10, devoted to the more general context of the sensitivity of
furanic polymers to degradation.

(72)
1246 A. GANDINI and M. N. BELGACEM

Table 1 collects the thermal properties of the various polyamides described in this work.
Among other relevant features, one interesting aspect arising from these data is the fact that
when the furan ring is inserted in the 2,5-mode, it introduces a higher degree of stiffness than
when it appears as 3,4-units.
An original approach to a fully furanic polyamide, consisting in the polycondensation
of N-hydroxymethylfuramide 75, was published in 1968.89 The authors claimed that they
obtained the same polymer as when using 5-carboxyethyl-2-furfurylamine 76, which
purportedly self-condensed liberating ethanol. The mechanism proposed for 75 is in fact
the same as that involved in the linear polycondensation of 7 discussed in Section 4.3,
namely, the loss of a water molecule between OH and H5, promoted by the electrophilic
activation of the primary alcohol function. The structure of the resulting polymer 77 was only
characterised by IR spectroscopy of rather poor quality and this issue remains therefore open
to verification.

(75) (76)

4.6. Furanic polyurethanes

The synthesis and characterisation of furanic polyurethanes is a polycondensation domain


that has been developed mostly in the last decade. The use of 9 or of oligomers from it as diols
for the preparation of polyurethanes based on aliphatic and aromatic diisocyanates to be used
as foams and elastomers has been described, albeit within a rather technological frame-
work. 90 In fact, the first thorough study in this field was based on the use of various difuranic
FURANS IN POLYMER CHEMISTRY 1247

diisocyanates 20in conjunction with aliphatic diols. 18Y91


Kinetic measurements involving the
furanic diisocyanates or MD1 and n-butanol showed that the aromatic diisocyanate was
somewhat more reactive than 20. The ensuing furanic-aliphatic polyurethanes 78 were
characterised by infra-red spectroscopy, DSC-TGA and dynamic mechanical properties.
The conclusion was that the replacement of aromatic diisocyanates by the difuranic structures
gave materiials with similar features.

(78)

A comprehensive investigation on: (1) the synthesis of both types of furanic monomer, (2)
their reactivity in the formation of model mono- and diurethanes, (3) the properties of these
simple products, (4) the preparation of a large variety of polyurethanes, (5) the kinetics of
these polyc’ondensations and (6) the painstaking characterisation of all materials has been
carried out in our laboratory over the last several years. 19*92-95 This very thorough approach
was deemed necessary in view of the very scanty information available, not only on furanic
polyuretbanes, but also on the more fundamental aspects related to the reactivity and stability
of the various structures used as reagents or obtained as products. Moreover, it was important
to verify the validity of odd claims% concerning the possibility of obtaining polyurethanes
from monofunctional furanic alcohols like 7. The monomers and monofunctional model
compounds used in this systematic endeavour included both furanic and difuranic mono-
and diisocyanates (28,29,33,34 and 20) and furanic alcohols and diols (5,7,8,9,48,49,79,
80, 81,82and 83).

OH
OH
(801
1248 A. GANDINI and M. N. BELGACEM

5~IyO&oy~~56 6' 6' 6 ’


0 0

Fig. 15. ‘H-NMR spectrum of diurethane 86. (Reproduced by qtrtnission of Elsevier


Science Ltd. Copyright 1993. Reprinted from .)

Conventional aliphatic and aromatic alcohols, isocyanates, diols and diisocyanates were also
used both for model studies and in the preparation of “mixed” polyurethanes.
The kinetic study related to the synthesis of tens of furanic mono- and diurethanes’9*92,93
revealed the excellent stability of all these furanic-aliphatic (e.g. 84), furanic-aromatic (e.g. 85)
and all-furanic (e.g. 86) models, except for structures in which the urethane moiety was directly
attached to the heterocycle. Fig. 15 gives a typical example of the clear-cut structures obtained
with all model compounds. As for reactivity, the alcohols displayed strong steric effects, whereas
isocyanates were mote sensitive to electronic factors, as shown by some representative second-
order rate constants given in Table 2. The extremely high reactivity of 28 and 33, attributed to
the conjugation of the NC0 function with the (die&) fur-an ring, coupled with the lack of
stability of the corresponding urethanes, ruled out any interest in using 33 as a viable monomer.

/\ o:,,
n, K 0
0
(84)
FURANS IN POLYMER CHEMISTRY 1249

100 J I I I I 1 8

--C Conversion (W)

RO
+ (Ao-A)/A

0
0 20 I

40 60 80 100 120

Time (mn)

Fig. 16. Conversion and second-order plot for the reaction of diphenylmethane diisocya-
nate (:MDI) with diol 9 in DMF at 23°C. (Reproduced by permission of the American
Chemical Society. Copyright 1993. Reprinted from “.)
1250 A. GANDINI and M. N. BELGACEM

Table 2. Second-order rate constants for the reactions of formation of mono- and diurethanes and linear
polyurethanes

Alcohols and diols Isocyanates* Ref.


Ph-NC0 29
93
n-BuOH 37.6 75.2
93
Ph-CH*OH 8.3 21.4
93
7 5.0 9.4
93
8 7.1 15.2
48 3.2 - 93

93
48 0.5
79 29.1 - 93

93
5 9.5 18.0
42 0.5 - 93

9 56.2 - 92

sot 2.1710.33 - 92

81 1.0 92

Diisocyanates Alcohols* Ref.


7 nBu-OH
MD1 3.5 - 92
2,6-‘ID1 2.0 - 92
HMDI 0.9 - 92
20b 9.2 - 92
2oc 9.4 60.4 92

Diols Diisocyanatesf Ref.


2oc MD1
1,4-Butanediol - 80 94
Ethylene glycol 7.3 18.6 94
9 5.5 155 94
80 6.4 - 94
81 1.7 0.4 94
82 6.2 30.5 94

*Reactions conducted in 1,2-dichlorobenzene with [NCOlo = [OH], = 0.5 M at 110°C.


tThe two rate constants correspond to the reactivity of primary and secondary OH groups of diol80.
SReactions carried out in DMF with [NCO],, = [OH], = 0.2 M at 90°C for 2Oc and 23°C for MDI.

Polymerisations94s95 also involved a large variety of combinations of difunctional mono-


mers and led again to furanic-aliphatic, furanic-aromatic and entirely furanic polyurethanes.
Moreover, the heterocycle was placed as a side group or in the chain backbone, or in both
positions. These polycondensation reactions were conducted in DMAc either without a
catalyst at 70°C or with nBu&Lau2. Their kinetics94 were followed by FTIR spectroscopy
and gave good second-order plots as shown in the typical diagram in Fig. 16. Polymers were
recovered by precipitation in methanol and their characterisation included the standard spec-
troscopies, molecular weight determination, DSC and TGA and dynamic mechanical proper-
ties.94,95 All structures were verified and the expected units clearly identified, without
detectable defaults. Figs 17 and 18 portray examples of the spectra obtained for all the
FURANS IN POLYMER CHEMISTRY 1251

IO 3000 2000 1500 1000 500

Wavenumber (cm-l)

Fig. 17. FTIR spectrum of fully furanic polyurethane 93.

polyurethanes. The viscosity-average molecular weights were in the range 5000 to 20 000 and
I, values were close to 2 as expected for linear polycondensations conducted under stoichio-
metric conditions.
Most polymers crystallised and all could be annealed to their glassy state. The melting
and glass-transition temperatures depended of course on the specific structures, with aromatic
co-moieties obviously giving stiffer chains than aliphatic ones. Fig. 19 illustrates these
features. An instructive comparison comes from the four furanic-aromatic polyurethanes
87-90, in which MD1 was used in all instances and the furanic diol was varied. The Tg and T,,,
of each polymer is given below. The role of the mode of enchainment of the heterocycle(s) is
well reflected in the changes of these parameters.

(87) Tg = 97 “C, Tm = 207 “C.


1252 A. GANDINI and M. N. BELGACEM

(88) Tg = 45 “C, Tm = 175 “C.

(89) Tg = 40 “C, Tm = 185 “C.

(90) Tg = 74 OC, Tm = 168 “C.

Similar trends were observed among structures derived from a furanic diisocyanate and
various diols (e.g. 91 and 92) or with all all-furanic polymers (e.g. 93 and 94).

(91) Tg = 50 "C.
FtJRANS IN POLYMER CHEMISTRY 1253

(93) Tg = 55 “C, Tm = 180 “C.

P+
3
0
K
0
H

I4

(94)
I\
0
I\
0

Tg = 20 OC, Tm = 17O’C.
IO
K
0 -I- n

The thermal degradation of these polyurethanes started systematically around 220°C but the
subsequent behaviour in the TGA diagrams showed marked differences which pointed to a
strong correlation between the proportion of furanic moieties in the polymer and the extent of
carbonaceous residue, as shown in Fig. 20.
A further set of furanic polyurethanes95 was based on the concept of thermoplastic
elastomers, i.e. was composed of structures built from a diol (1 ,Cbutanediol, 9, 81 or 82),
a diisocyanate (MD1 or 20) and a flexible macrodiol, namely a poly(tetramethylene oxide)
glycol (PTMG), a poly(caprolactone) glycol (PCL) or a polybutadiene glycol (PB), all with
molecular weights around 2000. After having ascertained the correctness of the structures
obtained and the adequate results with respect to their DP values, this study concentrated
on two major properties of the corresponding polyurethanes: their thermal transitions and
their dynamic mechanical behaviour. Some selected polymer structures are given below
as 95-97.
1254 A. GANDINI and M. N. BELGACEM

1 6’

n -I

z
c3
& --%x.-c
8
I

I 5 4 3 2
\

6 (ppm)

Fig. 18. ‘H-NMR spectrum of polyurethane 92.

Fig. 19. DCS trace of the polyurethane obtained by polycondensation of 9 with hexam-
ethylene diisocyanate.
FURANS IN POLYMER CHEMISTRY 1255

100

SO

all-furanic

furanic-aromatic

20

0
0 100 200 300 400 500 600 700 800

T (“C)

Fig. 20. Comparative TGA thermograms under N2 of polyurethanes of different structures


(sweep rate 10°C min-‘). (Reproduced by permission of the American Chemical Society.
Copyright 1993. Reprinted from 94.)

(95)
1256 A. GANDINI and M. N. BELGACEM

(96)

R= ,,(&$&
H3C a3
(97)

The typical feature of these materials is of course the coexistence of different micro-
phases, which display both crystalline and amorphous characters. This is clearly reflected
by the observation of both r, and T,,, with different relative structural contributions. It is
precisely in the temperature range encompassed between these two transitions that the
polymers behave like crosslinked elastomers. The Tgs obtained in the present investiga-
tion95 were systematically below - 15°C and mostly around - 60°C. As for T,, values ranged
from 150 to 190°C except in the case of PCL-based structures. The viscoelastic properties
showed an adequate rubbery plateau with moduli around 10’ Pa. The elongation at break at
room temperature was typically around 500%. Fig. 21 shows a typical set of dynamic
mechanical plots related to different structures.
Another recent report on furanic polyurethanes deals with the reaction of 20 bearing a HC-
CH3 bridge with the OH groups from the various components of wood flakes in order to
compare its adhesive capabilities with those of commercial MDI.97 The results obtained in
this rather empirical piece of work indicated that the furanic diisocyanate was less efficient
than MDI, albeit only slightly.

4.7. Furanic polyschiff bases

The possibility of synthesising polyschiff bases containing furan rings was first applied in a
short study of the reaction of 5chloromethylfuraldehyde and the self-condensation product of
PURANSIN POLYMERCHEMISTRY 1257

-1.6 -

-2.5 + I I I I I I I
-200 -1.50 -100 -50 0 50 100 150 200
T (“C)

Fig. 2 1. Dynamic mechanical behaviour of sequenced polyurethanes made with constant


molar ratio (l/2/3) of PTMGkhain extender/MDI; chain extender: (a) 80, (b) 82 and (c&B.
(Reproduced by permission of Elsevier Science Ltd. Copyright 1995. Reprinted from .)

5 (98) with aromatic diamines.98 These polymers were insoluble in all solvents tested and
were only characterised by nitrogen analyses which were not in good agreement with the
postulated structures. They started decomposing between 220 and 300°C.

(98)
1258 A. GANDINI and M. N. BELGACEM

In another investigation, aromatic and aliphatic diamines were condensed with 5,5’-thio-
bisfurfural.99 Again the polymers were insoluble, except in formic acid. Unfortunately, no
clear-cut structural analysis of these products was reported and the only set of data concerning
them, e.g. 99, related to their electronic conductivity which was consistently very low, around
lo-” S cm-’ at room temperature.

A series of polyschiff bases prepared from 30 and different aliphatic, aromatic and
polymeric diamines were synthesised and characterised in recent investigations which
involved first the study of the corresponding model compounds. lao~lol The first structure
aimed at attaining a fully conjugated macromolecule derived from 30 and 1,4-phenylene
diamine. For this, both the reactions of 30 with phenylamine and of 4 with the aromatic
mono- and diamine were studied and optimised. The resulting trimeric molecules 100 and
101were a useful source of structural information and were therefore thoroughly inspected by
spectroscopic and other analytical means. loo The preparation of the corresponding polymer
102 was marred by precipitation of the product during the early stages of the polycondensa-
tion (DP values around lo), whether different solvents were tried or bulk conditions were
applied.
FURANS IN POLYMER CHEMISTRY 1259

These materials displayed, however, a significant degree of conjugation since they absorbed
in the visible with a A,,, around 470 nm and a conductivity of about 10d8 S cm-’ for the
undoped polymer and about lOA S cm-’ after doping with iodine. They were insoluble in all
common solvents and depolymerised by acidolysis in formic acid. They did not show any
melting (D!SC and optical microscopy) up to 350°C when they began degrading.
A similar, but poorly conjugated polymer (103) was obtained100 by replacing 30 with 98.
Here the conjugation was in fact limited to the aromatic ring and the two furan heterocycles
adjoining it through the imino functions. This polymer was soluble in numerous polar
solvents and showed a lower thermal stability with decomposition beginning around
260°C. Other polyschiff bases from 30 were non-conjugated materials made from hexam-
ethylene diamine and the furanic diamine 32 as complementary monomers. ‘O”The first of
these polymers (104) was a crystalline product possessing a regular structure with molecular
weight of up to 2500 and a melting point around 150°C whereas the second (105) showed
irregularities probably arising from side reactions provoked by 32 (difficulties related to
keeping its purity at a high level) and a certain chemical fragility associated with the methy-
lene groups attached to the furan ring.
r

-“-“x7+
\ / \ / \

0 0 NN N

0
n
(103)
1260 A. GANDINI and M. N. BELGACEM

In a second investigation, lo’ 30 was condensed with hydrazine in order to obtain another
conjugated polymer (106) based on a furanic polyazine structure. Here too, the various model
compounds corresponding to this polymer were first synthesised and characterised before
proceeding to the study of the polycondensation itself, which gave the best results when m-
cresol was used as solvent. The resulting materials possessed the expected structure 106 with
molecular weights reaching 2000 and a thermal decomposition starting at 270°C without any
sign of melting. The low-DP fractions, soluble in methylene chloride, gave a X,,, at 440 nm.
Their conductivity after doping with iodine ranged from 10m5to 10S4 S cm-’ at room tem-
perature, indicating semiconducting properties, as in the case of the aromatic/furanic counter-
parts 102 discussed above.

(106)

The insoluble and infusible character of the conjugated polyshiff bases 102 and 106 make
them intractable. Attempts were made to synthesise more amenable structures, in terms of the
possibility of solution or melt processing, by introducing soft segments in the form of (1)
oligo(ethylene oxide) moieties within the main chain or (2) aliphatic moieties as side
groups. lo1 The first approach could not be realised directly because, if one mixes hydrazine
and an oligo(ethylene oxide) diamine (Jeffamine, M, = 600) with the furanic diadehyde 30,
only hydrazine condenses with it. It was therefore necessary to prepare two “trimeric pre-
cursors” arising respectively from the condensation of 30 with an excess of Jeffamine and
from the condensation of an excess of 30 with hydrazine. These two macromonomers were
then mixed in stoichiometric conditions and yielded a plastic-like soluble polymer with a
molecular weight of about 20 000. to1This product was therefore made up of short conjugated
sequences like 106 alternated with short poly(ethylene oxide) blocks (106a) and could
represent an interesting material possessing an aptitude to both electronic and ionic
conduction.

The second approach to softening the rigid polyschiff base 106 consisted in preparing a
furanic diketone with long aliphatic chains and condensing it with hydrazine. lo’ The strategy
seems to work, but the synthesis of both these new furanic monomers and the corresponding
polyschiff bases 106b must be optimised.
FWRANS IN POLYMER CHEMISTRY 1261

(106b)

It has been reported that the condensation of 4 with aromatic diamines’02V’03 yields poly-
mers bearing either imino groups arising from the specific reaction of fur-fur-al with two
aromatic amino groups, or triene structures formed by complete ring opening. These claims
were not convincingly corroborated by exhaustive structural analyses and must therefore
await confirmation by a more systematic study.

4.8. Furanic polyimidazoles

The reaction of disubstituted 1,2,4,6-tetraaminobenzenes with 31b followed by vacuum


dehydration at high temperature yields furanic polybenzimidazoles 107 with molecular
weights of about 10 000. lo4

These polymers were found to be insoluble in most solvents and only sparingly soluble in
protic acids. Their thermal behaviour was in tune with the expected heat-resistant structures,
in that they did not show any decomposition below 4OO”C, as shown in Fig. 22.
1262 A. GANDINI and M. N. BELGACEM

200 400 600 800 1000

T (“C)

Fig. 22. TGA thermogram of polyimidazole 107b:---,in air and -, under nitrogen.
(Reproduced by permission of Htithig & Wepf Verlag, Base]. Copyright 1982. Reprinted
from ‘04.)

a
6

i;
FURANS IN POLYMER CHEMISTRY 1263

4.9. Furanic polyethers

The condensation of 9 with dibenzylic halides has been studied with typical phase-transfer
systems. lo5 The furanic diol was dissolved in concentrated aqueous NaOH containing a
tetrabutylammonium salt. This solution was mixed with a solution of dibenzyl chloride in
an appropriate non-miscible solvent and stirred at 65°C for several hours. The ensuing solid
yellowish furanic polyethers 108were precipitated in methanol from the organic phase and
character&d. Fig. 23 gives the ‘H-NMR spectrum of one of these polymers. Molecular
weights were around a few thousand and the spectroscopic analysis confirmed that the
polycondensation had indeed led to structure 108. These polymers melted at about 110°C
and had a glass-transition temperature close to - 10°C.

4.10. Miscellaneous systems


Poly[2,5-bis(dimethylsilyl)furan] 109 has been prepared from 1 by silylation of the corre-
sponding 2,5dilithiofuran, free-radical chlorination and treatment of the ensuing dimethyl-
chlorosilane derivative with metallic sodium. lo6

0I\
0
- nBuLi
(CH,),HSiCl

This polymer was expected to show properties arising from the conjugation of the furan ring
with the Si-Si bonds, which have electronic properties similar to those of C=C moieties.
However, its UV absorption remained at a low wavelength (250 nm) even for relatively high
molecular weights (about 10000). The photolysis of this polymer produced a progressive
decrease in M,, down to about 2000, with a shift in its A,, to 238 nm. Thereafter the
degradation was much slower, probably because of this hypsochromic shift which reduced
the photolytic activity because of the lower rate of photon absorption.
The furanic dialdehyde 30 was treated with the reductive complex TiC14-Zn in dry THF at
0°C for 10 h and gave a polymer corresponding to structure 110with DP, = 43. lo7Unlike its
1264 A. GANDINI and M. N. BELGACEM

aromatic analogue, this polydiol could not be converted into the corresponding poly(2,5-
furylene vinylene).

q---& TQ$
0 0 PVL (110)
The reaction of 9 with diethylene glycol bis-chloroformiate under phase-transfer condi-
tions similar to those described above for polyethers lo5 gave an orange elastomeric polymer
(111)with a Tg of 9°C. This is the first published report of a furanic polycarbonate.

The furanic dialdehyde 30 has also been used as monomer for polyacetalysation with
pentaerithritol in order to prepare the quasi-Spiro polyacetal 112. lo8 Before synthesising this
material, the corresponding trimeric model compounds were prepared so as to character&e
the relevant structures and optimise the experimental conditions. This preliminary study
confirmed the feasibility of the reactions which were therefore transposed to the actual
macromolecular synthesis. The resulting polymers, soluble in hexafluoro-2-propanol, but
insoluble in common solvents, gave spectroscopic and elemental analyses in agreement with
structure 112. Fig. 24 gives the ‘H-NMR spectrum of the corresponding model which shows
that the spatial configuration of the cyclic acetyl units induces a complex resonance pattern
for the corresponding methylene protons. The thermal decomposition threshold for this
polymer occurred at about 350°C.

Conclusions
4.11.
The field of polycondensation reactions with furanic monomers is quite young, as shown
by the fact that most of the systems for which good-quality investigations have appeared are
recent additions to the scientific literature. Future contributions are badly needed, not only in
terms of kinetics, mechanisms and structures, but also (and even more) as far as properties
FURANS IN POLYMERCHEMISTRY 1265

I I L I
7 6 5 4

6 (mm)

Fig. 24. ‘H-NMR spectrum of the model compound obtained by the reaction of 4 with
pentaerythritol.

and possible applications are concerned. The fact that most monomers for these stepwise
reactions are derived from 5 should also constitute a stimulus for moving towards the
industrial production of this first-generation furan bifunctional derivative, let alone its
potential interest as precursor of fine chemicals.

5. CHAIN POLYMERISATIONS

This section treats the various systems in which furanic monomers have been submitted to
polyaddition reactions and describes the properties of the polymers and copolymers formed
therefrom. The subdivision follows the standard practice based on the nature of the active
species involved, namely, free radicals, cations, anions and active sites on coordination
catalysts.
In each section we shall deal with two distinct approaches: on the one hand, the use of
furan and furan derivatives which polymerise through the intervention of the heterocycle
itself, and, on the other, the use of furanic monomers which bear a polymerisable function
external to he ring. In the former context, the polymer loses its furanic character. In fact, the
macromolecules are formed to the detriment of its original conjugated system, although in
most cases the heterocyclic structure is maintained because of the formation of di- and/or
tetrahydrofuranic sequences. Instead, in the latter context, the furanic heterocycles are main-
tained as such and appear in the polymer as moieties pendant to a classical, e.g. vinylic,
backbones. Side reactions related to these lateral furan rings can however also occur, as
underlined in the specific systems discussed below.
1266 A. GANDINI and M. N. BELGACEM

5.1. Radical polymerisation


The behaviour of furanic monomers in chain reactions initiated by free radicals can be
rather complex and differs at times considerably from that encountered with more conven-
tional alkenyl structures. lS4*Therefore, before tackling the actual polymerisation systems it is
necessary to examine the course of model reactions involving free radicals generated from
classical sources and simple furan derivatives. The addition of the primary radical occurs
predominantly at C2 or C5, with a kinetic preference for the unsubstituted position. The
resulting fury1 radicals 113aand 113bdiffer of course in structure, but display a common
sluggishness towards propagation with further furanic substrate molecules because of their
relatively good stabilisation.

RA-3
1 \ + x’
(113a)
o
,.---
C.
;
0
X
R
(113b)

The two major reaction pathways open to these intermediates are therefore: (1) coupling with
other primary radicals, or among themselves, to give the various possible mono- and bicyclic
dihydrofurans, like 114aand 114b;or (2) homolytic aromatic substitution resulting in the
corresponding furanic structure, like 115.Thus,for example, 1,27

X
(113a) (114a)

.-.x
l-2
2::‘.
0 R
(113b) (114b)

(Ph),C’ +
R
FURANS IN POLYMER CHEMISTRY 1267

and of course similar mechanistic patterns leading to 2,3- and mixed dihydrofuranic compounds.
This behavicur has been observed with a variety of primary radicals generated by thermo-
lysis, photobysis or radiolysis; namely, alkoxy, ‘ow’~ benzoyl, “I phenyl, ’ l2 hydroxy, ‘I3 and
many others.27V’14*115 Hydrogen-atom abstraction from furan by a primary free radical is
kinetically very unlikely with respect to addition, e.g. less than 5% with t-butoxy radicals. ‘16
The absence of oligomeric products beyond the mono- and dicyclic dihydrofurans 114
clearly indicates that the furylic radicals 113a and 113b are unable to propagate in the
classical mode of radical polymerisation because of their resonance stabilisation. This is a
very important point in terms of both the reactivity of furanic monomers in radical poly-
additions and the role of furan derivatives as perturbing agents in free-radical chain reactions.
In fact, kinetic and thermodynamic factors can give rise to features ranging from “normal”
behaviour to lack of reactivity, depending on the furanic monomer or comonomer used, and
going from little effect to strong inhibition, as a function of the structure of both the monomer
and the furanic derivative added to it. The following sections below illustrate these points
through a series of typical investigations.
If furan is incapable of homopolymerising in the presence of free radicals, its copolymer-
isation with a number of typical monomers has been shown to occur and, of course, the
corresponding mechanisms and products have aroused considerable interest.

5.1.1. The copolymerisation of furan and 2-methylfiran with maleic anhydride

Maleic anhydride (116) readily reacts with a variety of furan derivatives according to the
well-known Diels-Alder cycloaddition mechanism. This important reaction is discussed in
detail in Section 8, where its application to various facets related to the synthesis and
modification of furan-based polymers is dealt with in detail. The relevant point here is the
fact that it is possible to induce the radical polymerisation of an equimolar mixture of 1 (or 10,
or 2,5_dimethylfuran) and 116, as first recognised in 1970.‘17 The major issue at stake, since
this report has to do with the actual mode of propagation, can be formulated as follows: is the
real monomer the Diels-Alder adduct 117, as claimed by some authors, I’* or is it the 1:1
donor/acceptor complex 118, as proposed by others ’ 19-‘2’?

c ’c\o
0

,+o-
.
0

(118)
1268 A. GANDINI and M. N. BELGACEM

Very ingenious mechanistic and structural studies “9-‘2’ point quite convincingly to the fact
that above about 70°C the adduct 117 reverts to its constituents and cannot therefore be the
precursor of the polymer. In these conditions the predominant macromolecular structure is
that arising from the activation of the charge-transfer complex 118, namely 119:

The same applies to the copolymerisation of 10 with 116, but the polymers obtained have a
more complicated structure containing different isomeric units. 12’
However, the adduct 117 can in fact be polymerised by using initiators capable of giving
free radicals at about 50°C and the ensuing polymer’20 bears essentially the completely
saturated structure 120:

3%
0

0’0 0
(120)

It is interesting to note that certain saturated moieties are also encountered in the copolymers
prepared at higher temperatures, indicating that, although the major mode of propagation
involves the complex 118, adduct 117 also seems to participate, albeit modestly, in the chain
growth. This point is still a source of controversy because in one study, ‘I9 carried out at 20°C
using y radiation, the polymer displayed essentially structure 119.
These perfectly alternating copolymers have molecular weights of 1000-4000 depending
on the conditions used for their synthesis. Regrettably, little has been published about other
properties or possible applications of these materials.
A kinetic study of the copolymerisation of 116 with 1 using thioglycolic acid as initiator
has also been carried out.‘22
2,3-Dimethylmaleic anhydride does not copolymerise with 1 under the same general con-
ditions used for 116. I21

5.1.2. Copolymerisation of furans with acrylonitrile and other acrylic monomers

The earliest report on the possibility of copolymerising a furan derivative with acrylonitrile
(121) concerned the use of 10. ‘23The undiluted mixture of monomers was irradiated at room
temperature with a y-ray source. A mixture of crosslinked and soluble (M, 3500) products
FUR.08 IN POLYMER CHEMISTRY 1269

60 80

Volume fraction of furan (%)

Fig. 25. The effect of varying the volume fraction (in the feed) of 1 (0) and of 10 (U) on
the yield of the copolymer at 25°C. (Reproduced by permission 25 Htithig & Wepf Verlag,
Basel. Copyright 1992. Reprinted from .)

were obtained and characterised only by infra-red spectroscopy. Copolymerisation in differ-


ent solvents provided more tractable materials, albeit with lower DP values. Apart from
evidence of the presence of both comonomers units in the products, this study did not provide
any clear information about the mechanisms or the insertion mode of 10. Some 20 years later
this topic was revived in a more systematic fashion.‘24*‘25 The copolymerisations, initiated
either thermally or photolytically with AIBN, involved 121 on the one hand and 1,lO and 2,5-
dimethylfuran on the other. The first striking observation, common to all these systems, was
the very low yields obtained, typically less than 5% under conditions which usually give total
conversion with conventional monomers, including 121 alone. A clear relationship was
established between the relative amount of furan comonomer in the feed and the yield of
copolymer, as shown in Fig. 25, suggesting a very important retarding role of the hetero-
cycle. ‘25 Moreover, the DP values were only of the order of a few tens, which indicates the
occurrence Iof chain-breaking events caused by the furanic comonomer. A thorough NMR
analysis of these products showed the incorporation of up to about one-third of furanic units
when a tenfold excess of that comonomer was used in the feed. These units (122-125) were
either in the form of 2,5- and 3,bdihydrofuran rings or as saturated cycles, the latter being the
more prominent the higher the proportion of 121 in the feed. ‘24,‘25Of course, the copolymer
structure wals completed by the dominant (minimum two-thirds) standard 126 unit from 121:

CN CN CN
(‘121) (122) (123)
1270 A. GANDINI and M. N. BELGACEM

NC

&
CN

CN
0

(124)
R CN
(125)

CN
(126)

These systems were modified by the addition of a Lewis acid like ZnCl*, which brought about
a clear improvement in the rate of copolymerisation and the DP values, although some
crosslinking occurred. lz5 These effects were most marked with 1 and 10. Aluminium-based
Lewis acids acted likewise, but less efficiently. Sulfur dioxide inhibited the copolymerisa-
tion. ‘25 The mechanistic origin of the positive and negative effects of the various additives
were not analysed in detail.
The copolymers were stable up to 250-300°C and left about 50% of carbonaceous residue
when heated to 1100°C in a nitrogen atmosphere. 126
Other furan derivatives were used as comonomers with 121, namely, 7, 9, 50 and bis-
(acetoxymethyl)furan. 126The reaction conditions were the same as with the simple furans and
the products had DP values around 100. The insertion of the furanic comonomer took place in
the same way as discussed above for furan and alkylfurans. Yields remained low, but reached
values of around 50% in some cases, with partial precipitation when the proportion of 121 in
the feed was high. The chemical modifications occurring when some of these copolymers
were pyrolysed or treated with strong acids were followed by solid-state NMR and it was
shown that extensive aromatisation took place. ‘26 Furfural and 5 were found to inhibit the
radical polymerisation of 121. ‘26
The use of a strongly electron-accepting monomer like methyl-a-cyanoacrylate (127) as
comonomer for 1 was tested in a photo-induced radical initiation reaction carried out in bulk
at room temperature with AIBN. ‘*’ Conversions ranging from 20 to 50% were obtained as the
proportion of 127 in the feed was increased and the corresponding mole fraction of dihy-
drofuranic units in the copolymer decreased from 45 to 27%. These units were predominantly
in the 2,5-configuration and the copolymer richer in furan had therefore a near-alternating
structure close to 128: ‘27
- CN

0
0
0 0
\ 0’
+ ’ n
CN P?
(127) (128)
FURANS IN POLYMER CHEMISTRY 1271

Furan has also been copolymerised with trisubstituted ethylenes bearing acrylic and cya-
noacrylic groups. These free-radical reactions were carried out in bulk or in acetonitrile
solution using AIBN. 128The resulting products were only characterised by elemental analysis
and solution viscosity. One of them was, however, also inspected by 13C-NMRspectroscopy,
but no clear evidence of its structure was provided.
All the systems above show that simple furans not bearing an external polymerisable
moiety are incapable of homopolymerising and are not good comonomers for radical
copolymerisation. This behaviour is related to the resonance stabilisation of the fury1 radicals
113aand 113bwhich results in coupling or hydrogen-atom abstraction, rather than propaga-
tion. Only highly electron-donating comonomers like 116, 121 and 127 manage to give
copolymers through crosspropagation, albeit with modest rates and DP values. Of course, even
in those most favourable conditions, the furan comonomer never enters in excess of 50% (rfU= 0)
and the structures obtained with the maximum furan content correspond therefore to alternating
copolymers.
The following sections are devoted to the radical polymerisation and copolymerisation of
“real” furanic monomers, i.e. molecules with an unsaturation external to the heterocycle.

5.1.3. 2-Alkenylfurans
It is understandable that 21a, which is the exact furanic counterpart of styrene, should have
attracted considerable attention since the very beginning of polymer science applied to
radical reactions. A large number of patents, issued between the two World Wars, were
devoted to its polymerisation and to the uses of the ensuing products.’ More scientific
investigations followed’ in various countries to show that free radicals can initiate the
bulk, 129,130 solution I31and emulsion 132S133polymerisation of 21a to give the corresponding
linear polymer 129. Its thermal polymerisation in benzene indicated a termolecular initiation
and produced the Diels-Alder dimer 130 in a side reaction. 134

-c J
.

‘CT 0'

(1291 (130)

Table 3. Reactivity ratios for various radical copolymerisations of 21a (MI) with different vinylic
compounds (M2)

M2 rl r2 Ref.
Styrene 1.9 0.25 130

Acrylonitrile (121) 0.82 0.037 130


Vinylidene chloride 11.7 0.15 136

Butadiene 4.52 0.11 137


1272 A. GANDINI and M. N. BELGACEM

The y-irradiation of bulk 21a leads to a crosslinked product”5 to which was assigned the
following structure:

The problem with the more traditional radical systems involving 21a is the fact that
polymerisations are sluggish and give low values of DP when compared with the behaviour
of conventional monomers, like styrene, treated in the same conditions. Only emulsion
polymerisations were claimed to give high yields and very respectable molecular
weights. ‘32~133The reactivity ratios published for the radical copolymerisation of 21a (M,)
with various monomers 130~136~‘37 are given in Table 3.
More recent work27V’38,‘3ghas shown that in systems other than emulsions, 21a indeed gave
regular, low-DP structures like 129, but, as the polymer accumulated, it began to play the role
of a radical trap through its pendant unconjugated furan rings and yields were drastically
reduced as a consequence. The mechanism of this termination reaction can be written as:

~+Q_#&+
(131) (129)
- (132)

or as the corresponding reaction of 129 with primary radicals.


Emulsion systems on the other hand, gave excellent results, I39 both for homo- and
copolymerisations. The optimisation of these polymerisations provided quantitative yields
of high-molecular-weight materials. 13’ The reason for the major difference in phenomenol-
ogy between homogeneous and emulsion media is unclear, but could be related to the unique
feature of emulsion polymerisation which induces the presence of only one free radical in the
active micelles. ’
Another serious problem related to both homo- and copolymers of 21a is their high
FURANS IN POLYMER CHEMISTRY 1273

sensitivity tlo oxidation, which is certainly related to the mobility of the hydrogen atom
attached to ,the tertiary carbon atom (see 129) and which leads to crosslinking when these
materials are exposed to the atmosphere. 1~139 This very important aspect will be discussed in
the more general framework of Section 9.
A careful NMR analysis of 129 has been reported. 14’
21b behaves like its unmethylated parent, but gives higher yields and DP values. 14’ The
presence of the methyl group at C5 obviously slows down the termination reaction (see 131+
129 above) between macroradicals (or primary radicals) and pendant heterocycles.
2-Isopropenylfurans 21c and 21d do not homopolymerise by free-radical mechan-
isms, 142~143as in the case of their aromatic counterparts, but give copolymers with several
142.144
comonomers.

5.1.4. Furjiuyl acrylate and methacrylate

A series of pioneering studies on 25b showed first’45 that its radical polymerisation in
solution proceeded smoothly up to high conversions and only in the very last stages did some
crosslinking occur. Its bulk polymerisation’46 followed all the typical features displayed by
acrylic monomers associated with the Trommsdorff effect. Its copolymerisations with styr-
ene 147~148 and methyl methacrylate 149were found to give reactivity ratios in good agreement
with the relative ease with which each comonomer can delocalise the unpaired electron, i.e.
values not far from unity. In other words, no specific problem related to the presence of the
furan heterocycle seemed to arise in these systems. This conclusion can be rationalised by the
observation of the structure of 25 which would give rise to radical chain carriers 133 char-
acterised by a reasonable stabilisation due to the neighbouring ester group, just like with all
conventional acrylates and methacrylates:

In other words, this macroradical would be expected to react predominantly with the acrylic
unsaturation of the incoming monomer, rather than with its furan ring moiety. Normal
propagation would thus be the rule and, addition to the heterocycle, the exception. It is
therefore reasonable that high yields of high-DP polymers and copolymers are obtained with
these monomers and the intervention of side reactions leading to crosslinking are only
observed if the systems are pushed to extreme situations.
The polymers and copolymers of 25 have been the object of several recent studies. A
homopolymer 134 with an M, of 85 000 was prepared by standard radical initiation and
submitted to TGA analyses in the temperature range lOO-600°C under inert and oxidative
atmospheres. 15’
1274 A. GANDINI and M. N. BELGACEM

Its thermal degradation started at about 250°C and led to partial depolymerisation accom-
panied by intramolecular cyclisation between neighbouring groups, with the expulsion of
furfuryl alcohol, and intermolecular crosslinking through oxidative coupling at C5.
The free-radical copolymerisation of 25a (M i) with 2-hydroxyethyl methacrylate 15*and of
25b (M i) with N-vinylpyn-olidone 152~153 have been studied in DMF. The first system151
gave rl = 0.9 and r2 = 1.3 and the Ts of the various copolymers showed the expected variation
between 50°C for the furanic homopolymer 134 and 90°C for the other. The second
system 152J53gave rl around 4 and r2 around 0.004, pointing to the very limited
stabilisation of the macroradical ending with a pyrrolidone moiety. Similar variations of
the T, of the corresponding random copolymers were observed, compared with those of
the other copolymerisation system. I51Figs 26 and 27 provide pictorial evidence for these
trends.
These materials were submitted to thermal degradation, which was followed by TGA. 154
Whereas poly(N-vinylpyrrolidone) was stable up to 38O”C, all copolymers began to degrade
well below 300°C. The copolymers were also studied as hydrogels and their swelling beha-
viour followed gravimetrically. 155The hydrophilic character of the copolymers was essen-
tially attributed to the N-vinylpyrrolidone units. In a similar vein, hydrogels resulting from
the copolymerisation of 25a and hydroxyethyl methacrylate were assessed in terms of
swelling kinetics. 156The obvious interest of all these hydrogels is in applications related to
biocompatibility.

5.1.5. 2-Vinyl&routes
Contrary to furfuryl acrylates and methacrylates, the potential monomers 24 show a
remarkable inertia towards radical polymerisation. Forty years ago various authors had
already reported that 24a did not homopolymerise with free-radical initiators157*158and
indeed retarded strongly the polymerisation of vinyl acetate. 159Later, a more specific study
was conducted on the oligomerisation of 24a by AIBN. 160The authors of this study found
again that with a “normal” initiator concentration, no polymerisation occurred and only
when this was raised to 15-20% of the monomer concentration did they isolate 10 to 20% of
oligomers with an M, of 700. These products, claimed to have the “official” vinylic structure
142a, had in fact an NMR spectrum showing a very high resonance contribution from the
initiator fragments and anomalous integration ratios for the various protons attributed to
the monomeric units. In particular, a large peak between 4 and 6 ppm clearly suggested
FURANS J.NPOLYMER CHEMISTRY 1275

’ (a)
10.8 -

0.2 0.4 0.6 0.8 1.0

F, (feed)

1.0

- (b)
/
0.8 - ,
,
/
.
/ ;
/
/
,
0.6 _ , b
/
,
5 / .A
/
,
/
0.4 _ II .~..‘,... Q
/
/
/
/ **.e....“’
/
. ...“’
/’ . . ..
0.2 _ / . ..*
,
/ . . ..d
/ . . ...*
, *....... **
/ /...... .*a’
/
._...” d.” I I I I
0 0.2 0.4 0.6 0.8 1.0

FP

Fig. 286.Composition diagrams of the (a) furfmyl acrylate/2-hydroxyethyl methacrylate


and (b) furfuryl methacrylate/N-vinylpyrrolidone copolymer systems. (Reproduced by
permission of Elsevier Science Ltd and John Wiley & Sons, Inc. Copyright 1993.
Reprinted from 15’~152.)

the presence of a large proportion of dihydrofuranic and vinylic moieties. Indeed, a more
careful investigation27 of the reaction between AIBN and 24a showed that these oligomers
(M, 600 to 900) are in fact a complex mixture of units including an average of two 2-cyano-2-
propyl end groups, 2,5- and 4,5dihydrofuranic rings (the former being in larger amounts),
vinyl groups and 2-substituted furan heterocycles. This pattern suggests that the primary
1276 A. GANDINI and M. N. BELGACEM

. (4
. -. -.
-\
350

300 I 1 1 1

0 0.2 0.4 0.6 0.8

WF

(b)
400

cs
Do
b 350

300
0 0.2 0.4 0.6 0.8 1.0

WF

Fig. 27. Glass-transition temperatures of the (a) furfuryl acrylate/2-hydroxyethyl metha-


crylate and (b) furfuryl methacrylate/N-vinylpyrrolidone copolymer systems as a function
of furfuryl acrylate and furfuryl methacrylate weight fractions, respectively, in the
copolymer chain. (Reproduced by permission of Elsevier Science Ltd and John Wiley
& Sons, Inc. Copyright 1993. Reprinted from ‘5’,152.)

radical, produced by AIBN (135), as well as oligomeric radicals derived from 24a (136),
added onto the monomer, either through the normal propagating route (i.e. onto the vinylic
double bond) to give the classical radical 137, or at the C5 position of the furan ring, thus
giving stabilised furylic radicals 138.These then generated the dihydrofuranic moieties by
the coupling reactions discussed above. The two modes of addition occurring here can be
therefore summarised as follows:
FURANS IN POLYMER CHEMISTRY 1277

R.’ + I \
0 oL.H
%
0

(138)

(135) (1361

A similar behaviour was observed” with 2-vinylfuryl acetate (139), 2-vinylfuryl propio-
nate (140) and Z-vinylfuryl acrylate (141). However, 24b polymerised in the presence of high
concentrations of AmN to give low-molecular-weight products with the expected regular
structure 142b with no detectable evidence of the presence of vinylic or dihydrofuran
groups. l7

*
5- I‘“i
‘I” 5 0

(14Za)
0

n
0

0’

(142b)
-
-Cl

n
1278 A. GANDINI and M. N. BELGACEM

The latter system suggests that the C5 position is less likely to be the site of addition of
primary and oligomeric radicals when it bears a substituent and thus the normal propagation
mechanism can proceed with less-frequent side reactions. However, the size of the substituent
at C5 does not seem to play an important role since 24c gave the same features 16’as 24b,
whereas higher DP values had been expected on account of the larger steric hindrance of the
t-butyl group which, in principle, should have shielded more effectively the C5 position from
radical attack.

5.1.6. Miscellaneous furanic monomers


2-Fury1 vinyl ketone29 (26) and 2-fury1 oxiranei6* (Wa) respond very meekly to free-
radical initiation and give oligomers without any real interest. In a different context, a recent
report ‘63 describes the pyrolysis of 5-methyl-2-furfuryl benzoate (143) at 650°C and the
vapour-deposition polymerisation of the resulting 2,5-dimethylene-2,5dihydrofuran (144)
to give the novel furanic polymer 145, which was described as a semicrystalline material with
Tp = 20°C and T, = 110°C. This colourless product was insoluble in several solvents and
started decomposing around 385°C. More work is needed on this interesting system, parti-
cularly to verify the structure postulated for the polymer and to explain its insoluble character.

cc
a
‘00
-
0
(143)
0 &k \
(1441

5.1.7. Retardating and inhibiting role of furans

The fact that the furan heterocycle should be considered as a free-radical trap is quite
convincingly documented by the numerous reports discussed in the preceding sections. In-
deed, simple furan derivatives do not homopolymerise despite their dienic character precisely
because the initiation step leads to radical structures that are too sluggish to propagate
because of their resonance stabilisation. *’ Moreover, with furanic monomers, i.e. structures
bearing the furan ring and a potentially polymerisable external unsaturation, three types of
behaviour have been observed, namely: (1) inhibition, as with 24a, due to the higher reac-
tivity of the ring compared with that of the vinylic moiety;27S160
(2) practically normal poly-
merisation and copolymerisation behaviour, as with 25, because the free radical resulting from
addition to the external C=C double bond is stabilised by the neighbouring groups; ‘45-‘56and
FURANS IN POLYMER CHEMISTRY 1279

(3) retardation, as with 21a and 21b, because both the ring and the alkenyl moieties can be
attacked and also because the dead polymer molecules which accumulate in the medium can
also react with free radicals.129-‘41To these observations one must add the fact that simple
furans, used as comonomers in systems involving conventional monomers, slow down very
considerably the rates of polymerisation and give very modest yields even with high initiator
concentratians. Evidence also exists, as already pointed out, to show that when the C5 position
of the ring in alkenyl furanic monomers is substituted, its vulnerability to attack by free radicals
is decreased., which is to the advantage of “normal” propagation and structures.
Several studies have been conducted to investigate the role of furanic compounds, added as
simple furans or as comonomers, in free-radical polymerisation systems. When the structure
of the furaric additive is highly conjugated, as with 2-furfurylidene derivatives (146), the
and can reach activities similar to those of typical com-
inhibition is very efficient27’161’165
mercial inhibitors like hydroquinone. Here, the reaction of a free radical with such a “tri-
enic” system, coupled with a carbonyl or a CN function, leads to a very highly stabilised
addition product 147, which of course will not propagate at all. Furfural has a similar action
as recently confirmed by a study in which traces of this compound in 21a resulted in very poor
rates and yields in emulsion systems. 139

&, R= 8-R' or &N

(146)

(147)

When simple furan derivatives are added to typical radical polymerisation systems, retard-
ing effects ;are always detected, albeit of a very different magnitude, depending on the furanic
additive and the monomer used.166-173Thus, for example, with a typical AIBN-initiated
solution polymerisation of vinyl acetate, i.e. a monomer giving a poorly stabilised free
radical, concentrations as low as 4 x 1O-3M of 10 already produce retardation and
the presence of 20% v/v of the same additive inhibits the polymerisation. 166~167~169-173 25
Dimethylfuran was found to be less active, confirming the barrier effect of double o-sub-
stitution. 0ther compounds, such as furoic acids, gave stronger effects. Thus, the ratio
between the values of the propagation and the degradative transfer (addition to the furan
ring) rate oonstants, were 0.02 and 0.23 for 10 and 15, respectively. The corresponding ratios
between the values of the rate constants related to initiation (135 plus vinyl acetate) and
addition to the furan additive (135 plus furan derivative) were 1200 and 500, respec-
tively. 166~169-173At the other end of the spectrum, the polymerisation of methyl methacrylate
was not affected by the addition of fairly large amounts of 10 or 2-ethyl furoate, and indeed,
1280 A. GANDINI and M. N. BELGACEM

even with up to 50% v/v of the former and 20% v/v of the latter, this system behaved
“normally”* ‘66,168With methyl acrylate, the retarding effect appeared with concentrations
of 10around 0.1 M (rate of propagation halved) and inhibition occurred with 50% v/v. 166~168
Finally, with styrene, retardation only begun with about 20% v/v of 2-ethyl furoate and the
polymerisation was inhibited when the concentration of this furan derivative was raised to
50% v/v. 166*168
The third context refers to copolymerisation systems involving a conventional monomer
and a furanic comonomer used in 50/50 v/v conditions. With 24a, the use of styrene resulted
in complete inhibition, the use of vinyl chloride gave retardation and the use of methyl
methacrylate produced the homopolymer of the latter. 16~,16’This is in tune with the relative
stability of the different macroradicals involved. The copolymerisation reactivity ratios pub-
lished for 21a with various comonomers (Table 3) point to the same tendencies. Finally, 25b
was found to copolymerise “normally” with methyl methacrylate and styrene ‘@ as well as
with hydroxyethyl methacrylate, 15’ as indicated by reactivity ratios close to unity. 147-149,151
A recent detailed study of the homopolymerisation of 25a and 25b showed’74 that some
crosslinking occurs in these systems, suggesting that even a supposedly trouble-free furanic
monomer like 25b gives some side reactions involving the addition of macroradicals and the
C5 position of heterocycles present on dead polymer chains.

5.2. Cationic polymerisation

The use of acidic initiators to promote the polymerisation of furan derivatives must again
distinguish between two general situations, namely, whether the monomer is a simple furanic
derivative (including furan itself), or a molecule containing both the heterocycle and an
external polymerisable function. Moreover, the dual dienic/aromatic character of furan lends
itself to electrophilic substitution reactions which are highly regiospecific. This feature can be
both problematic and useful, depending on the specific context, as discussed below.

5.2.1. Furan and its alkyl derivatives

2-Methylfuran 10 was the first of simple alkylfurans to be submitted to acid-catalysed


polymerisations in the early 1960s. After several investigations175-‘79 involving the use of
Lewis acids in dry media at - 30 to 6O”C, the most likely structure proposed’77,178 involved
mainly 2,3-dihydro-2-methylfuran units (148):

PL
0'

n
(148)

However, the experimental evidence brought forward to corroborate the validity of this
structure was too scanty to constitute an acceptable proof. Indeed, subsequent studies with
both Lewis and Bronsted acids indicated a more complex situation in which the oligomers
formed contained both furan and dihydrofuran rings, but also OH groups and aliphatic
In particular, the treatment of 10 with dilute phosphoric acid gave tetramer
moieties. 180V’81
FURANS IN POLYMER CHEMISTRY 1281

149 resulting from the partial acid-catalysed hydrolysis of one furan ring coupled with
condensation reactions.

(149)
2,5_Dimethylfuran is less prone to oligomerisation ‘75than 10, but the reasons for this lack of
reactivity have not been explored.
Furan (1) is quite sensitive to acidic media and yields dark-coloured products which can go
all the way to black crosslinked resins if drastic conditions or long reaction times are used.
The first study of the mechanism of this system, which is still one of the best, involved strong
organic Bronsted acids in the dry bulk medium. Is2It was shown that some of the repeat units
of these polymers were dihydrofurylene structures which resembled those obtained from the
analogous treatment of cyclopentadiene, i.e. long conjugated sequences responsible for the
strong coloration and the electronic conductivity of the readily protonated materials. A later
investigation called upon a closer inspection of the structures obtained. lg3It was found that
trichloroacetic acid led to soluble polymers, whereas trifluoroacetic acid gave a mixture of
soluble and crosslinked products and perchloric acid led to totally insoluble resins. The DP
values of the soluble polymers ranged from 50 to 300. The mechanism postulated on the basis
of the two studies invokes the protonation of 1 at C2 to give the delocalised carbenium ion
150. The ensuing propagation involves the electrophilic addition of this species onto the C2
position of the monomer, which leads initially to the dihydrofuranic macrocation 151.Sub-
sequently, however, this linear structure undergoes branching and eventually crosslinking
through the intervention of the residual double bonds on the dihydrofuranic rings. The final
structure would therefore resemble formula 152, which contains both substituted dihydro-
and tetrahydrofuran heterocycles. These conclusions were corroborated by NMR and IR
spectra and the knowledge of the cationic polymerisation mechanism of 2,3-dihydrofumn.‘W
Fig. 28 gives a typical ‘H-NMR spectrum of polyfuran obtained by cationic initiation.
1282 A. GANDINI and M. N. BELGACEM

&+f-
0 0'

o/ o/ 0

(152)
The end groups of these polymers consisted of the three types of ring. The presence of large
proportions of furan rings in these structures were explained in terms of a high degree of
branching and therefore of terminal groups. Finally, the colour of the polyhnans could not
be rationalised in terms of structures 151and 152, which are not conjugated Dihydrofuran ring-
opening was thus postulated and structures like 153 advocated to justify the deep colour of the
polymerising solution and indeed of polymers obtained at high conversions. The reaction of 1
with acidic surfaces in the presence or absence of gaseous Lewis acids and with or without y-
irradiation has also been studied, albeit in much less detaiLnu

An interesting study of the oligomerisation of 1 by aqueous HCl gave two major products:
tetramers 154 and 155. M The mechanisms proposed for their respective formation are given
below:

0’ 0
+


-Q-tJ$9
4
0 0 -
O/
3

b ‘\
O
oh
d \ I\
0 ’
O I\
% 0 /
0
FURANS IN POLYMER CHEMISTRY 1283

(154)

3-
0' <OH
/w
\/0

0’
-
+
0
’ etc
-

(155)

These routes are reasonable, but it is quite surprising that no acid-catalysed hydrolysis of the
furan ring was invoked in this context. Both this and the previous studies on the homogeneous
action of acids on 1,‘82~183~186
were taken up again a few years later. 187With aqueous HCl, the
authors remarked that indeed the two tetramers were formed (with 155being in excess), but
tended to tnm into resinous products at long reaction times and the typical black intractable
tar was the final issue of these systems. Treatment of 1 with trichloroacetic acid gave
again a soluble mixture of oligomers in which both tetramers were claimed to be present
1284 A. GANDINI and M. N. BELGACEM

8 6 4 2 0

~(PP~)

Fig. 28. ‘H-NMR spectrum of polyfuran. (Reproduced by permission of John Wiley &
Sons, Inc. Copyright 1980. Reprinted from 18’.)

in the ether-soluble fraction, on the basis of spectroscopic and chromatographic evidence,


with 155 dominating. A pentamer was also detected by mass spectrometry. The rest of
the product, soluble in chloroform, was composed of higher oligomers, which suggested
the presence of units similar to those found in both tetramers, but again structures like
155 seemed to be much more abundant. The DP, of these products was about 10 and its
DP, close to 50. In conclusion, this work proposed a basic mechanism for the acid-
catalysed polymerisation of 1 similar to the set of reactions sketched above, but with a
predominance of addition reactions induced by the protonated furan ring 150 with the
positive charge at C3. The presence of a carbonyl band in the IR spectrum of the
polymer obtained with aqueous HCl clearly suggests some hydrolytic ring opening of
furan rings, a fact which was not evoked in the publication. Curiously also, no mention
was made to the origin of the intense colour of these polymers,‘87 despite the obvious fact
that all the structures proposed would give by definition colourless products (unconjugated
substituted furan rings absorb below 250 nm).

5.2.2. 2-Alkenyljiuans

Structures 21 are highly nucleophilic and therefore particularly adapted to cationic initia-
tion. Early reports on this type of system L88-1soreflected qualitative work which just pointed
FURANS IN POLYMER CHEMISTRY 1285

to the polymer&ability of 21a and to the complex structure of the coloured products obtained.
The first attempt to look into the kinetic and mechanistic features of one such system involved
21a with iodine as initiator and methylene chloride as solvent, in the temperature range 20 to
50°C. 19’External orders of one were found with respect to both monomer and iodine and the
structure of the deeply coloured oligomers obtained, inferred from an unpublished IR
spectrum, ‘was claimed to be the simple linear architecture 129. A later brief study on this
same system clearly showed that, in fact, these products had more complicated structures
which closely resembled those obtained with trifluoroacetic acid, as discussed in detail
below. 192
A comprehensive investigation of the cationic polymerisation of the four monomers 21
was undertaken a few years later with the same solvent, but with (mostly) trifluoroacetic acid
(TFA) as initiator and covering a wide range of temperatures between - 78 and 40°C.20*192-199
Reactions were conducted in an inert atmosphere with highly purified components. The
kinetics of polymerisation were followed dilatometrically in VCICUO, but also by measuring
the monomer consumption by ‘H-NMR. The changes in the electronic spectrum and in
electrical conductivity were monitored with a special high-vacuum device provided with
both a quartz cell and platinum electrodes.20 Moreover, a number of model compounds were
prepared and tested in order to unravel certain mechanistic problems. 193~195~‘98 All four
monomers readily polymerised even under the mildest conditions and in this respect they
resembled ‘other very nucleophilic monomers like vinyl ethers. The initiation with TFA led to
the formation of both the corresponding esters and the related carbenium ions. Whether the
propagation was pseudo-cationic or truly cationic depended on the specific conditions, but all
mechanistic evidence favoured the former mode in most instances in which therefore ionic
species were virtually absent.
Before discussing detailed features, it is instructive to outline the general phenomenology
of the polymerisation of 21a at room temperature, since it will serve as background for all
further discussion related to the behaviour of each of the four monomers in terms of kinetics,
mechanisms and structures.20’192*193

1. All polymerisations went to completion.


2. Titration of the acid at any stage of the reaction (including long periods after the comple-
tion of the polymerisation) gave values identical to the initial concentration used.
3. Treatment of the reaction mixture with solid potassium hydrogen carbonate produced the
total removal of the acid.
4. Addition of a second portion of monomer at the end of a given polymerisation produced
further polymerisation, albeit at a lower rate.
5. Polymerisation solutions turned from pale yellow at low conversion to an amber-like
colour at later stages to dark brown towards the end of the reaction. The colour continued
to deepen even after complete conversion, reaching dark blue-grey tonalities or even
blackish1colours when the solutions were kept under vacuum for weeks.
6. The electrical conductivity of the polymerising solutions grew steadily as the reaction
proceedled and kept growing even after complete consumption of the monomer had been
reached.
7. Neutralisation of these deeply-coloured conducting solutions induced both a sudden par-
tial bleaching and a drop in conductivity, leaving a light orange-brown, poorly conducting
organic phase. Re-acidification brought back the original colour and high conductivity,
1286 A. GANDINI and M. N. BELGACEM

and the magnitude of these two parameters was higher the stronger the new acid solution.
This cycle could be repeated indefinitely.
8. If an empty side arm of the reaction vessel was cooled during the polymerisation, a certain
amount of each of the three components vacuum-distilled into it, as shown by the onset of
a new, albeit slower, polymerisation accompanied by colour formation, etc.
9. The yellow-brown polymers obtained in these systems were in fact oligomers, with DP
values ranging from 5 to 20, which displayed large structural irregularities with respect to
the normal vinylic chain 129.

The other 2-alkenylfurans (21b-d) gave similar phenomenologies, but with the specific
absence of some of the features enumerated above, as will be described in detail in the
following paragraphs.
All evidence gathered indicated clearly that the growth of these polymer chains followed at
least in part the classical vinylic propagation course, but that this normal route was accom-
panied by side reactions whose high relevance here was associated with the specific chem-
istry of the furan heterocycle.
The first of these side events concerned the ease of electrophilic substitution at C5 when
this position was free. ‘92-‘95This phenomenon was detected when certain polymerisations
were followed directly by ‘H-NMR spectroscopy, for the relative intensity of the H5 reso-
nance decreased progressively when compared with those of H3 and H4, which remained
stable and identical between themselves. Thus, with 21a and 21c, the linear chain growth was
marred by this type of interaction which took place on furan rings belonging to both monomer
and polymer chains, as shown in the following mechanisms which, applied to 21c, gave
therefore the two branched structures 157 and 158. (The active species 156 are shown as
ion pairs for simplicity, but, as stated above, in most situations they were in fact the corre-
sponding polarised trifluoroacetate esters.)

+& - @-@&-%A
(1561 (1571

With 21a, around room temperature, the same interactions occurred, but they were accom-
panied by the anti-Markovnikoff addition of the active species 159 to the C5 position which
gave structures 160 and 161:
FURANS IN POLYMERCHEMISTRY 1287

I I I I I I I I I
8 7 6 5 4 3 2 1 0

6 (wm)

Fig. 29. 'H-NMRspectrum of 162. (Reproduced by permission of Htithig & Wepf Verlag,
Basel. Copyright 1983. Reprinted from 195.)

&.@+y42
j - B-
(159) (160)

At lower temperatures, the latter alkylation reaction became predominant, not only with
respect to ihe “normal” one, but also with respect to propagation and the polymer prepared
1288 A. GANDINI and M. N. BELGACEM

around - 20°C had a structure made up almost exclusively of 162. Fig. 29 illustrates this
point through iH-NMR spectroscopy.

All these structural anomalies were assessed by very detailed inspections of the NMR and IR
spectra of the products. Moreover, apart from the careful spectroscopic analyses of these
polymers and of the polymerising solutions as a function of the extent of monomer consump-
tion, model compounds were also used to confirm all the mechanistic features given above. 195
Thus, for example, the reaction of 21a with an excess of 10 at room temperature, 195carried
out under the same conditions as the corresponding polymerisation of 21a, gave the two
compounds 163 and 164, which clearly confirmed the occurrence of both types of electro-
philic substitution discussed for this monomer.

(163) (164)

Of course, monomers 21b and 21d did not give alkylations at C5 since that position was
substituted by a methyl group. Moreover, no evidence of a similar electrophilic reaction at C3
or C4 was ever detected.19’ The latter feature rejoins all the observations related to the
polycondensation of 7 discussed in detail in Section 4.3, namely that the regiospecificity
of this type of substitution is particularly high, with the C5 position being practically the
exclusive site of the event.
The alkylation reactions described above are not specific to furan rings, since they have
been observed in the cationic polymerisation of styrene and other vinylaromatic mono-
mers.200-202 What is peculiar to the present context are the quantitative kinetic features,
i.e. the much more relevant contribution of this type of reaction with the furanic monomers.
In fact, branching and eventually crosslinking with styrene becomes apparent mainly when
systems are left standing without neutralisation after the end of the polymerisation, since then
the macromolecular active species have the time to attack (preferably) the C4 positions and
thus build progressively the ensuing ramifications. As shown above, the same phenomena
occur here at the same time as propagation, since in part they constitute an alternative to it and
cannot therefore be avoided. In other words, with aromatic monomers, the alkylation reac-
tions are much less important and only involve active species plus polymer molecules,
whereas with alkenylfurans, they compete with propagation and concern both monomer
and dead polymer molecules. Apart from the trick of appending methyl groups to the C5
position of these furanic monomers, which constitutes a “cheating” solution, it was found
that alkylation can be reduced with 21c by reducing the polymerisation temperature because
the activation energy for this reaction is higher than that of propagation. 195,203Conversely, it
was found that lowering the dielectric constant of the polymerisation medium produced a
FURANS IN POLYMERCHEMISTRY 1289

(4
I
Cc)

(b

400 500 600

h (nm)

Fig. 130.Typical evolution of the UV-vis. spectrum during and after a “slow” polymer-
isation of 21a. [2la], = 0.70 M; [TEA],, = 6.0 x 10” M; T = 25°C; solvent = CH2C12.
-- 1 cm cell; - - -, 1 mm cell; (a) 150 min; (b) 25 h; (c) 17 days; (d) 42 days.
(Reproduced by permission of Hiithig & Wepf Verlag, Basel. Copyright 1982. Reprinted
from *O.)

corresponding decrease in the contribution of propagation through the alkenyl moiety,


relative to monomer alkylation. *03
The second side reaction occurring in parallel to the propagation of the cationic poly-
merisation of 2alkenylfurans concerns only interactions between active species and dead
polymer molecules.*’ The first evidence for the occurrence of these reactions was the appear-
ance of colour during the cationic polymerisation of 21a and 21b, but not with 21c or 21d.
The visible spectra of the coloured solutions became more intense and moved bathochromi-
tally as the reactions proceeded. Figs 30 and 31 give typical examples of this evolution
related to 21a and 21b, respectively. At the same time as the chromogenic events, the
electrical conductivity of the polymerising solutions increased progressively. Both the optical
density at a given wavelength and the conductivity displayed S-shaped curves when plotted
against reaction time as shown in Figs 32 and 33. Also, as stated above, both phenomena
continued to occur after complete monomer consumption. The neutralisation of the deeply
coloured solutions and their re-acidification followed the typical pattern illustrated by Fig. 34,
which clearly shows that the neutral polymers were themselves deeply coloured, albeit less
intensely than their protonated counterparts.*’
1290 A. GANDINI and M. N. BELGACEM

Fig. 31. Typical evolution of the UV-vis. spectrum during and after a polymerisation of
21b. [21b],, = 0.53 M; [TFA]o = 2.3 x 10” M up to 4 h, then second acid phial broken to
give a total [TFA],, of 7.7 x 10m3M; T = 25°C; solvent = CH$Z12. Spectra taken from a
1 cm cell for the first 2 h;, from a 1 mm cell thereafter. Times as indicated on trace.
(Reproduced by permission of Htithig &fr;;p;aVerlag, Base]. Copyright 1982. Reprinted
.)

0
0 50 100 150
Time (mn)
Fig. 32. Typical variation of the optical density at 480 nm during and after a polymerisa-
tion of 21a. [21a],, = 0.40 M; [TFAlo = 4.0 x 10e2M; T= 25°C; solvent = CH,C12; 1 cm
cell. (Reproduced by permission of Htithig t Wepf Verlag, Basel. Copyright 1982.
Reprinted from 20.)
FURANS IN POLYMER CHEMISTRY 1291

Time (mn)

Fig. 133.Typical evolution of the electrical conductivity during and after the polymerisa-
tion shown in Fig. 32. (Reproduced by permission of Hiithig & Wepf Verlag, Basel.
Copyright 1982. Reprinted from ‘O.)

The mechanisms reflecting these events must involve the formation of progressively longer
conjugated sequences through interactions not requiring the monomer, but implicating, as a key
qualitative factor, the hydrogen atom on the vinylic carbon attached to the ring. The presence of
polymer molecules bearing a terminal unsaturation (165) is also indispensable to these events.
Their formation reflects the classical monomer-transfer reaction from an active species 159:

(165)

The following set of reactions” explains adequately the phenomenology of the second side
reaction:
1292 A. GANDINI and M. N. BELGACEM

A + +-

B-

(159)
3-
0’

(165)

(166) (167a)

159

+c&p-
166
+ /fjy

(167b)

167b + B- - HB +

/y&A/Q etc*

(Mb)
FURANS IN POLYMER CHEMISTRY 1293

A (nm)

Fig. 134.Combined spectrum of the solution of the experiment shown in Fig. 5, after
opening the device under NP. Successive dilutions allowed to scan down to 300 nm in a
1 cm cell. (a) Just after opening (45 days from mixing the reagents); (b) after neutrahsation
(methylene chloride phase); (c) after re-acidification of the methylene chloride phase to
give I:TFA] = 3.0 x lo-’ M. (Reproduced by permission of Httthig & Wepf Verlag, Basel.
Copyright 1982. Reprinted from ‘O.)

At each cycle, one polymer molecule acquires a saturated end and the corresponding carbenium
ion 167 arising from this H- transfer becomes more conjugated than its precursor unsaturated
molecule. ‘The allylic structures 167 thus generated are all in equilibrium with their neutral
counterparts 168 bearing conjugated double bonds. Globally therefore, each new double bond
resulting from the successive loss of H- and H+ is accompanied by the corresponding conver-
sion of a classical active species into a dead end-saturated polymer molecule 166. This
sequence of events keeps taking place as long as the system is kept “alive” in vucuo, as shown
in Figs 30 imd 3 1. In some instances, blackish colours were attained because the solutions were
left for several months.20
1294 A. GANDINIand M. N. BELGACEM

240 280 340 420

A(nm)

O-

-I
440 600 700 900

A(nm)

Fig. 35. UV-vis. spectra of 129 isolated after a long-lasting ionogenic reaction. The
different regions were scanned by successive dilutions with dry methylene chloride.
(Reproduced by permission of Htithig & Wepf Verlag, Basel. Copyright 1982. Reprinted
from *O.)

The presence of polymer molecules bearing different numbers of conjugated double bonds
is also clearly visible through the multiple absorption shown in Fig. 35 and the actual values
could be inferred from model compounds like 169.20*204

As with the branching reactions arising from electrophilic substitutions, the growing of
multiple unsaturations is not unique to 2-alkenylfurans. Monomers like styrene*” and in-
FURANS IN POLYMER CHEMISTRY 1295

dene206give similar phenomenologies, but at much lower rates: with the former, the appear-
ance of these conjugated moieties occurs well after the end of the actual polymerisation,
whereas with the latter coloration begins during the polymerisation, but to a much lesser
extent than with 21a and 21~.
Concerning the two side reactions, it can be concluded that 21a was exposed to both and
therefore the number of different unit structures borne by the ensuing polymers prepared at
room temperature included, apart from 129, arising from the normal propagation step, the two
types of S-substituted branching and the multiple terminal unsaturations together with chains
ending witb saturated groups. Monomer 21b gave linear coloured products bearing the same
type of multiple unsaturations and saturated end groups as those from 21a. The isopropenyl
homologue: 21c gave colourless polymers possessing the ramifications shown by structures
157 and 158, except for the products prepared at - 78’C which were essentially linear.
Finally, at room temperature, 21d gave oligomers with the regular structures 170 which
displayed two types of unsaturated end group arising from transfer reactions. At - 78°C
the DP values of these linear colourless polymers reached about 100.

-
lY 0'

(170)
n

As mentioned above for 21a, no termination reaction was detected in these systems, except
for the formation of the allylic carbenium ions 167 formed with 21a and 21c, which were
incapable of adding to the monomer. Thus, the longer these particular systems were left after
complete polymerisation, the lower the availability of the acid for re-initiation upon a second
monomer addition. With 21b and 21d, the systems were “living” in the sense of no termina-
tion in the absence of basic impurities.
Transfer reactions were important with the four monomers. Thus, the DP values of the
product formed between room temperature and 40°C were limited to 2-10 and only below
about - 50°C could these materials be called polymers.
The kinetics of the cationic polymerisation of the four 2alkenylfurans were also thoroughly
studied. 1%~‘97~*993203
With monomers 21a and 21c, the formation of stabilised allylic ions like
167 was synonymous with acid consumption, i.e. progressive “internal termination”, which
resulted in a dampening effect on the rate of polymerisation. This decrease at each subsequent
monomer addition described above for the general phenomenology related to 21a is thus
readily explained, although when the monomer was added at the end of a given polymerisation,
the colour intensity of the solution decreased, indicating that a certain proportion of the allylic
ions 167 were deprotonated by the excess monomer, i.e. by sheer mass action. With monomer
21c the kinetics were less complicated, but the occurrence of the alkylation reactions still gave
fairly complex patterns which became much more tractable when the temperature was lowered
to - 78”C, because these side reactions were then practically eliminated.
1296 A. GANDINI and M. N. BELGACEM

The specific structure of the various active species responsible for propagation, transfer and
side reactions was also the object of scrutiny20”g3~‘g5~1g8~203 and, although ionic species play a
role in certain conditions, most mechanistic events seemed to arise from ester molecules
when trifuoroacetic acid was used as initiator. Model reactions with the carbinols related to
monomers 21 were useful in establishing that conclusion.
If, on the one hand, the multiple concomitant reactions related to 21a in an acidic environment
constitute a source of complication both in terms of structural anomalies and kinetic/mechanistic
behaviour, they can, on the other hand, be exploited to prepare specific materials. The useful
applications of the C5 electrophilic substitution will be discussed in Section 5.6 devoted to the
general context of the use of furanic additives to prepare original macromolecular structures.
The relevant aspect here is the polymerisation of 21a under drastic conditions, namely its
treatment with high concentrations of trifluoroacetic or methanesulfonic acids in bulk or in
concentrated solutions.207 In these aggressive media, the systems rapidly develop intense black-
ish colorations and crosslink to give hard brittle materials. Even after neutralisation, the result-
ing resins maintain a black colour and can be ground into fine carbon-like powders. The severe
acidic conditions were shown to induce electrophilic substitutions also at C3 and C4. The
properties sought for these products were related to the presence of long conjugated sequences
attached to insoluble networks, as shown by the schematic structures 171.‘07

1 \

3% 0'
-
0 /
0'

/ol
3
‘n
NO
-
(171)

As expected, they showed a remarkable aptitude to fix protons because of their strong Lewis-
base character; i.e.
nHA+C=C+HC-C+ +(HA),_,A-
where C=C stands for a conjugated sequence.
The proton-scavenging property of these polymers reached up to 60 meq g-’ which reflects
a very strong activity. *07The fact that the proton exchange occurred heterogeneously and that
the protonated resins could easily be regenerated by treatment with a strong base, opens the
way to applications such as purification of solvents containing acidic impurities and removal
of the acidic products in alkylation, acylation or similar reactions, without the complications
arising from the need of isolating a soluble protonated side-product, as with pyridine for
example. It is important to note that these resins only operate in organic media and lose their
FURANS IN POLYMER CHEMISTRY 1297

activity in aqueous solutions, i.e. the excess of water keeps the electrophile sufficiently
solvated and stops it from attacking the resin.

5.2.3. 2-Vinyl furoate

It has aheady been pointed out in Section 5.1.5 that this compound does not polymerise via
radical initiation because of its tendency to capture the free radicals through the furan ring. Its
cationic polymerisation with various Lewis acids ‘*‘08is a curious phenomenon in the sense that
vinyl esters are usually not activated by this type of initiation. With 24a the vinyl moiety is
protonated and the resulting carbenium ion 172 alkylates the C5 position of a second monomer
molecule. The succession of these events generates the polymeric structure 173 which is an
original polyester with little apparent connection to the structure 24a of its precursor:

‘%f
I\
0
0
o+
+
B-
+
0,
”0
0
Oe -

(172)

5.2.4. Furfural and homologues

The polymerisation of aldehydes through the carbonyl bond is marred by the fact that the
enthalpic gain associated with the propagation (change of the C=O a-bond into a C-O u-
bond) is quite modest and therefore these monomers often have low ceiling temperatures,
with the notable exception of formaldehyde.2w Conjugated aldehydes are in a particularly
unfavourable situation because this state of affairs is worsened by the electronic delocalisa-
tion in structures like C=C-C=O, as with acrolein. Thus, for example, benzaldehydes do not
polymerise”” even at very low temperatures. Both 4 and 6, but also 5-bromo-2-furaldehyde
and 2-fury1 methyl ketone, have been shown to behave likewise under a large variety of
conditions2’ ’ which included different types of initiator and temperatures down to - 78°C or
even condensation with a magnesium catalyst on a surface cooled at liquid nitrogen tem-
perature. In all instances, these potential monomers were recovered quantitatively, although
1298 A. GANDINI and M. N. BELGACEM

60

50

10

0
100

Time (h)

Fig. 36. Typical polymerisation curve for the reaction of 6 with methylstyrene (MS). The
arrow indicates the second initiator addition. T = 0°C; [6] = [MS] = 1.8 M; [BFE], =
[BFE], = 0.2 M. (Reproduced by permission of Elsevier Science Ltd. Copyright 1977.
Reprinted from *Il.)

of course some catalysts certainly reacted with the carbonyl moiety, i.e. “initiation” took
place.
The thermodynamic restrictions related to the lack of propagation with these systems can
be overcome if a second monomer is introduced which has a “normal” behaviour, i.e. a
ceiling temperature well above the reaction temperature associated with a high absolute value
of AH,. This trick obviously will only allow the insertion of at most 50% of the non-poly-
merisable aldehyde units in the copolymer structure, since it will remain impossible to have
sequences containing more than one successive element arising from the aldehyde. In other
words, the AH associated with these propagation steps will refer either to the homopolymer-
isation of the “normal” comonomer, or to an alternating step.
The application of this principle to furanic aldehydes 4 and 6 was successfu12”V30band gave
a number of interesting polyether-type copolymers resulting from the cationic copolymerisa-
tion with alkenyl and heterocyclic comonomers. The first study”’ concentrated upon alkenyl
partners and the search for optimal conditions related to the relative nucleophilicity of the two
comonomers; as expected, it turned out that the carbonyl function in 4 and 6 is a very basic
moiety and that therefore the best comonomers were highly nucleophilic molecules. The most
interesting system found in this investigation involved the use of 6 with cr-methylstyrene in
toluene at 0°C with boron fluoride etherate (BEE) as initiator, but other solvents and lower
temperatures gave similar results. Reactions showed an induction period and yields tended to
taper off around 40%, although the copolymerisations could be re-activated by adding more
FURANS IN POLYMER CHEMISTRY 1299

(a)

I
I I I 1 I I I I I
4000 3000 2000 2000 1800 1600 1400 1200 1000 800

Wavenumbers (cm-‘)

6 4

6 (wm)

Fig. 37. (a) FAIR and (b) ‘H-NMR (CDC13) spectra of a typical copolymer Oor-methyl-
styrene. (Reproduced by permission of Elsevier Science Ltd. Copyright 1977. Reprinted
from *‘I.)

initiator as shown in Fig. 36. The reactivity ratios were found to be 1.0 and 0 for 6 and
cr-methylstyrene, respectively. These data were quite surprising given the clear evidence
of complete lack of homopolymerisation of 6. However, the copolymers prepared with
molar fractions of 6 above 0.2 contained more than 50% of units from 6, which indeed
reached 0.74 with monomer feeds very rich in 6.211 Their structure, determined by IR and
NMR spectroscopy (see Fig. 37), was entirely compatible with the expected enchainment and
1300 A. GANDINI and M. N. BELGACEM

it was therefore concluded that a penultimate effect occurred in these copolymerisations


allowing the presence of sequences of two and even three furanic comonomer units.
On the other hand, a-methylstyrene was so much less nucleophilic than 6, that its
incorporation was limited to 50% at best, i.e. even when its contents in the feed were as
high as 80%. Thus, structure 174 shows a typical alternating copolymer which is in fact a
polyether and 175 gives an idea of the polyether-polyacetal obtained when the incorporation
of units from 6 exceeded 50%. The molecular weights of these copolymers were in the range
of several thousands.

0 0 0

~~
L
0
0’

(174)
-

Pa 0
o\o’
-

(175)
-
n

The copolymerisation between 4 and a-methylstyrene gave similar phenomenologies2”


but with two major differences: (1) the maximum incorporation of units from 4 did not exceed
60% and (2) some black resins accompanied the copolymer, particularly with high molar
fraction of 4 in the monomer feed. An additional problem was the occurrence of some
alkylation at the C5 position of the furan ring of the units of 4 inserted into the copolymers,
which gave rise to branched structures. The use of other comonomers with 4 was also studied.
Structures that were less nucleophilic than cz-methylstyrene, e.g. isobutene, isoprene, cyclo-
pentadiene and styrene, produced inhibition phenomena because the attack of the initiator
onto the carbonyl bond of 4 gave a cationic species which was too stabilised to react with the
comonomer. Indene instead behaved essentially like ar-methylstyrene, although both the
copolymer yields and the incorporation of units from 4 were smaller. Conversely, with
comonomers that were more nucleophilic than a-methylstyrene, like acenaphthylene and
N-vinylcarbazole, their near-homopolymerisation dominated with very little incorporation
of the furanic partner.2”
In a second study of the cationic copolymerisation of 6, different vinyl ethers (ethyl,
chloroethyl, isobutyl and 1,2-dimethoxyethylene) were used as comonomers in methylene
chloride at - 5 to 20°C using first BFE as initiator. 30bWith this class of highly nucleophilic
n,?r-donor molecules, the products showed again that both comonomers were inserted into the
macromolecule and also that up to about 70% of units from 6 could be introduced. The
copolymerisations were, however, considerably slower than the corresponding homopoly-
merisations of the vinyl ethers, indicating that the complex between the initiator and 6 was
not excessively reactive. Molecular weights were in the range of several thousands, not very
different than those obtained from the homologous homopolymerisations of the vinyl ethers.
All the copolymers displayed regular structures 176 which were in accord with the expected
mode of insertion of each comonomer, i.e.
FURANS IN POLYMER CHEMISTRY 1301

3-\ R'
0’

0'
R

R = ethyl, chloroethyl, isobutyl or methyl

R’ = H or methoxy
(176)

The use of other Lewis acids and in particular the organoaluminium chlorides gave very
similar qualitative features with differences in yields and other quantitative aspects which
were however rather modest. 30bIn these systems, the various donor sites in the copolymer
units showed a marked tendency to complex with the initiator, which resulted in a progressive
slowing down of the rate of monomer consumption and the attainment of limiting yields.
A new combination consisted in mixing 6 and tetrahydrofuran (11) and using methyl
triflate as imtiator. 30bAlternating copolymers were obtained only if 11 was added in excess,
because otherwise the complex between MeOTf and 6 was too stable. The spectral analysis
(see the ‘II-NMR spectrum in Fig. 38) confirmed that this polyacetal had the expected
structure 177, with an M, of about 5000. The use of triethyloxonium tetrafluoroborate as a
possible alternative initiator was unsuccessful because this salt formed a strong complex with
6 which did not attack 11.

(177)

In conclusion, 4 and 6 can be copolymerised with various alkenes and n-donor monomers,
provided the right balance is found in their relative basicity. In fact, if the comonomer is too
weakly nucleophilic, the complex between the initiator and the furanic aldehyde remains
inactive and no polymerisation or copolymerisation occurs. Conversely, if the comonomer is
too strongly nucleophilic, it shows a predominant tendency to homopolymerise, ignoring the
presence of the furanic aldehyde (although the latter might slow down the process). When
appropriate combinations are obtained, alternating copolymers possessing regular polyether
(or polyacetal in the case of 11)structures are readily obtained, albeit sometimes at modest
rates and yields. Moreover, in certain instances it is possible to introduce short sequences of
1302 A. GANDINI and M. N. BELGACEM

8 ; a ; 4 ; 2 1 0

6 (ppm)

Fig. 38. ‘H-NMR spectrum of a copolymer 6111.

furanic units in these copolymers thanks to penultimate effects, i.e. to synthesise materials
containing more than 50% of lateral heterocyclic moieties.

5.2.5. 2-Fu&yl vinyl ether

This monomer combines two structural features that contribute to its very pronounced
nucleophilic character, namely the furan ring and the vinyl ether moiety. It is therefore not
surprising that its cationic polymerisation occurs with great ease and at rates that are
extremely high compared with those of more conventional monomers, including the alkyl
vinyl ethers. 3obUnfortunately, with 27a the alkylation reactions at C5 also take place very
readily and the polymers have therefore a marked tendency to crosslink at the early stages of
the reactions. The 5-methyl homologue 27b has only been studied briefly, but a problem
related to the presence of traces of 8, coming from the transvinylation synthesis, left doubts
about the validity of the results obtained. 3obObviously, more work is needed to gain a better
grasp of the potential of these monomers.
FURANS IN POLYMER CHEMISTRY 1303

5.2.6. 2-Fqfurylidene methyl ketone and homologues

When 22a and 22h are heated in the presence of an acidic catalyst, they polymerise all the
(note: their reactions in basic media will be discussed in
way to b1ac.kinsoluble products 1,4*7,48
Section 5.3.6). These resins have found numerous applications as corrosion-resistant coat-
ings, abrasive surfaces and materials with good thermal stability. The mechanism(s) leading
to the linear polymeric structure, and subsequently to the branching which induces the final
crosslinking, are complex and the first investigations, aimed at understanding their details,
were not very convincing in terms of evidence.43@ Thus, various condensation reactions, e.g.
between the carbonyl group and the H5 atom, were proposed, as well as the direct interven-
tion of the furan ring which was claimed to be absent in the resulting polymers.
These hypotheses were, however, dismissed by later work which was based on a careful
elemental, UV, IR and NMR analyses of the oligomers (M, = 1000, Ts = 100°C) obtained
from bulk 22a treated for several hours at lOO-140°C with phosphoric or benzenesulfonic
acid. 43 The basic structure 45 obtained has already been given in Section 5.1.3, but it is
necessary here to justify its validity. First of all, these materials had elemental analyses
which were very close to that of the monomer, a fact that excludes the predominance of
condensation reactions. Second, in their IR spectra, all the typical bands characteristic of the
monosubstituted furan ring were present as in the monomer, which was taken as clear
evidence of’the minor (if any) intervention of the heterocycle in the polymerisation mechan-
ism. Moreover, two major changes were observed in the IR spectra: the decrease of the band
at 1630 cm-’ attributed to the C=C moiety and the appearance of a band at 1725 cm-’ which
indicated the formation of unconjugated C=O groups (the C=O in 22a absorbs at 1675 cm-’
and this band decreased in the spectrum of the polymer). Third, the UV spectrum of the

(b)

t
3oc

A (nm)

Fig. 39. UV-vis. spectra of (a) poly(22a) and (b) 22a.


1304 A. GANDINI and M. N. BELGACEM

Fig. 40. ‘H-NMR spectrum of a poly(22a).

polymer showed a decrease in the intensity of the major band at 313 nm resulting from the
Fu-CH=CH-C=O chromophore in the monomer and the appearance of a peak around
220 nm, typical of unconjugated furan compounds, plus a minor broad absorption extending
into the visible. Fig. 39 shows these differences.
Finally, the ‘H-NMR spectrum of these products (see Fig. 40) compared with that of its
precursor gave a ratio of 5/3 between aliphatic and furanic + olefinic protons instead of 3/5 for
22a. Additionally, the integration of H5 was close to l/2 that of H3 + H4. All this evidence43
corroborated the validity of the assumption that the main polymerisation mechanism in the
first phase of the acid-catalysed reaction of 22a is simply the propagation through its alkenyl
moiety, and that therefore structure 45 is a good approximation of the reality.
The light brown colour of these polymers and the fact that both C=C and conjugated C=O
groups are still present in their structure, although to a minor extent, suggested the presence of
unsaturated end groups like 178, i.e.

(178)
FURANS IN POLYMER CHEMISTRY

5000 3000 2000 1800 1400 1000 650

Wavenumbers (cm-‘)

Fig. 41. IR spectra of (a) 22h and (b) poly(22h).

At higher temperatures (200-500°C) and for longer reaction times, these products cross-
link according to complicated mechanisms which involve condensation reactions, as shown
by the relealse of water,*l* but also of CO, methane and C02.*13
The difuranic unsaturated ketone 22h was also studied in a similar context.44 The polymers
obtained (M, = 3 100 and Tg = 250°C) gave IR spectra in which a new band appeared around
2900 cm-’ (absent in the monomer). The other features in these spectra were the conservation
of the furanic character and the same changes in the C=C and C=O regions described above
for 22a. Fig,. 41 shows these comparisons. The electronic spectra also showed the partial loss
of conjugation, the appearance of the simple furanic absorption around 220 nm and the weak
broad band in the visible extending now beyond 600 run (here the monomer 22h already
absorbs up to 450 nm because of the high extent of conjugation). In conclusion, the behaviour
of 22h resembles very closely that of 22a, i.e. both are initially polymerised by a propagation
reaction involving the external C=C unsaturation(s). However, with 22h the presence of
two such moieties gives early branching which facilitates crosslinking even in the absence of
the condensation reactions occurring at high temperature. The two structural units present in
the polymer of 22h are therefore 179 and 180, apart from the unsaturated end groups similar
to 178.
1306

- A. GANDINI and M. N. BELGACEM

3
\O

O I

1.0 i 1n
(179) (180)

A more recent study on the thermal polymerisation (at 120- 180°C without added acids) of
22h,*14 as well as on the structural features of the oligomers formed by its acid-catalysed
thermal treatment (up to 160°C with phosphoric acid),215 proposed the intervention of
numerous reaction mechanisms including Diels-Alder cycloadditions, which were suspected
of being the culprit for the formation of dense crosslinked structures at later stages. Unfortu-
nately, the evidence put forward to sustain these hypotheses was not totally explicit and it
would have helped to report the spectral features concerning these materials.

5.2.7. Miscellaneous firanic monomers

2-Fury1 vinyl ketone 26 does not respond very vigorously to cationic activation.29 Thus,
with BFE in methylene chloride, yields were limited to about 15% of oligomeric products in
which the expected vinybc units 184were accompanied by other structures, mostly involving
the inclusion of the heterocycle in the main chain through alkylation reactions at C5. This
behaviour confirmed the high reactivity of the furan ring with respect to electrophilic sub-
stitution at C5, an aspect already repeatedly discussed above. The low tendency of 26 to
polymerise can be rationalised in terms of competition between the C=C and the C=0 sites
for the Lewis acid, i.e. between an initiation reaction and a complexation which is equivalent
to inhibition.
2-Furyloxirane (23a) was shown to polymerise in methylene chloride under the influence
of various cationic initiators. 162 The oxirane ring-opening propagation leading to linear
polymers was accompanied once again by extensive C5 alkylation of the furan heterocycle
and the systems tended therefore to give crosslinked materials. In order to avoid this com-
plication, the corresponding 5-methyl derivative 23b was synthesised and tested. The ensuing
polymers remained soluble in the reaction medium, but their structures showed evidence of
electrophilic substitutions at C3 and/or C4. This was confirmed by the fact that, if the
polymerisation systems were left for prolonged periods of time, gelation occurred. This is
the first example of a cationic polymerisation involving a furanic monomer in which the C3
and/or C4 positions were clearly shown to be susceptible to electrophilic attack in competi-
tion with the C5 counterpart, which however remained the predominantly reactive site. This
original behaviour must be related to the activating role of the oxirane moiety. In conclusion,
although monomers 23 have a good aptitude to cationic polymerisation, the ensuing pro-
cesses are marred by important substitution reactions on the furan ring, which are detrimental
FURANS IN POLYMER CHEMISTRY 1307

to the thermoplastic character of the materials obtained, i.e. branching and ultimately cross-
linking reactions cannot be avoided.

5.3. Anionic polymerisation

Up until a decade ago, reports on the anionic polymerisation of furanic monomers were
singularly fscarce and focused essentially on the oligomerisation of 22a. This paucity of
scientific activity probably stemmed from the fact that many nucleophilic media can provoke
the destruction of the furanic heterocycle rather than activate a polymerisable function
attached to it.6 Nevertheless, if appropriate conditions are applied to specific monomers,
perfectly reasonable results can be obtained, as discussed in this general section.

5.3.1. Furan and homologues

Whereas lithiation of the furan heterocycle under the appropriate conditions has been
shown clearly to lead, depending on the substrate and the excess of organolithium reagent,
to either the 2- or the 2,5-lithio derivatives,216 which are useful precursors for a variety of
synthetic procedures, other metallation reactions claiming to give oligomers have no real
foundation for lack of serious characterisation of the products.1327217 The reaction of furan
with sodium naphthalene gives some resinification, but no 2-fury1 sodium.*i8 Sodium has
been reported to react with furan in an argon matrix at 4 K under visible light excitation to
give the radical anion 181 arising from ring opening.*19

H H

(1811

This species would obviously be extremely reactive in a liquid medium and attack the
unreacted furan. The so-called oligomers arising from this attack have not been identified,
but would very probably be a complex mixture of open structures and substituted rings. In
other words, unless mild lithiation conditions are used for synthetic purposes, the reaction of
strong nucleophiles with the furan heterocycle leads to ring opening and uncontrolled addi-
tion reactions, giving a mixture of unidentified low-molecular-weight products. This is also
true of 2-alkylfurans.

5.3.2. 2-Alkenylfurans

It was first claimed 40 years ago that 21a was polymerised by butyl lithium. **’ A few years
later, it was shown that the reaction of metallic sodium with the same monomer only gave
unidentified oligomers, but no anionic vinylic polymerisation, although the reaction product
was active for the initiation of the anionic polymerisation of acrylic monomers. **’ A further
1308 A. GANDINI and M. N. BELGACEM

report insisted on the polymerisability of 21a with sodium naphthalene and even on the living
character of the poly(2-vinylfuran) anion, which was claimed to initiate the polymerisation of
methyl methacrylate. 132S2’7 These assertions were totally deprived of any proof related to
initiation, characterisation of the active species or indeed polymer yield or analysis. The
claimed block copolymer was described likewise, without the least characterisation. 132*217
This optimistic description was challenged later with a brief mention to the fact that 2-
alkenylfurans do not respond to sodium naphthalene activation.***
This puzzling state of affairs was cleared by a series of systematic experiments conducted
under rigorous vacuum conditions with thoroughly purified and dried reagents and sol-
vents. 223The results of this study can be summarised as follows.

n-Butyllithium does not induce the polymerisation of 2alkenyl furans 22a, 22b and 22c at
room temperature. This is true in THF and hexane and with initiator concentrations as high
as 2 M. No polymer was detected for several days. 2 to 5% of monomer was consumed to
give an orange resin whose IR and NMR spectra bore no relation with the expected
furanvinylic structure 129.
The same monomers were not polymerised by a sodium mirror in THF between - 20 and
30°C. Again a resinification reaction took place, but the coloured product only revealed the
presence of complex conjugated and branched oligomers resulting from partial cleavage
of the heterocycle. The spectra of these products had no resemblance to those of the
expected polymers.
The same monomers mixed in THF with sodium naphthalene (whose concentration was as
high as lo-* M) in the temperature range - 78 to 30°C produced the instantaneous
disappearance of the green colour of the initiator which was slowly replaced by a pink
shade. However, no polymerisation occurred within several days. More than 95% of the
monomers was recovered and the residue was again a resin possessing a complex structure
bearing no relationship to the expected structures.
The system 22a, sodium biphenyl in HMTP at room temperature gave a low yield of real
white poly(2-vinylfuran) (129), with the correct spectra and a DP, close to 100. Despite
several attempts to optimise the conditions, the limiting polymer yield never exceeded
about 10%.

These experiments223 proved unambiguously that the classical strong anionic initiators do
react with 2-alkenylfurans, but that the product of that initiation, probably the anion radical
182, is very rapidly destroyed by side events and propagation has little chance of occurring,
except when a specific system like 4 above is used.

(182)

However, even then, the active species did not survive long enough to ensure a high monomer
conversion. The fragility of the furan heterocycle in this context must be related to its dienic
character, since of course styrene responds very readily to these initiators and even 2-vi-
nylthiophene polymerises to 100% conversion with sodium naphthalene,222 although the
propagation rate with this monomer is some 100 times lower than that for styrene. Thus
FURANS IN POLYMER CHEMISTRY 1309

one finds hlere the typical correlation of the extent of aromatic/dienic character mentioned in
the Introduction, namely that thiophene has an intermediate electronic configuration between
those of blenzene and furan, which translates itself into the corresponding intermediate
behaviour in the anionic polymerisation of its vinylic derivative compared with styrene
and 21a.

5.3.3. Fur$wal

The thermodynamic impossibility of homopolymerising 4 has been discussed in Section


5.2.4. A single mention is found in the literature concerning the anionic copolymerisation of 4
and the comonomer used was dimethylketene.!” This system produced a stereoregular
structure, and the material displayed a sharp melting point characteristic of a highly crystal-
line polymer, indicating that the two comonomers had been inserted in a regularly alternating
fashion to give the polyester 183:

-
Pit 0
0

01

(183)
II

Regrettably, this original piece of work, carried out more than 30 years ago, has not been
followed b:y any further investigation on similar systems involving furanic aldehydes.

5.3.4. 2-Fury1 vinyl ketone

This monomer (26)* has attracted very little interest and only one study of its polymeri-
sability has been published. 29 Whereas free-radical and cationic initiation proved rather
inefficient in terms of yields (5-15%) and molecular weights (700-2000), anionic systems
gave reasonable monomer conversions, attaining 75%. However, oligomerisations (M, =
1000) in these nucleophilic media took place with the intervention of several modes of
propagation which included the normal vinylic activation, but also the intervention of the
furan ring and of the carbonyl function. Thus the three units 184,185 and 186 were detected
in these macromolecules:

(184) (1851 (186)

*WARNING! The toxicity of 2629 must be emphasised to warn chemists who might be interested in studying it.
1310 A. GANDINI and M. N. BELGACEM

::
;J

\O/cH*-NyC\N-cH2*
o- I I I
,H\ /A,
i I
CHZ

c
L

Y (cm-‘)

ii

I”.‘) ““““I.‘.‘l”.

7 6 5 4 3

6 &v-O

Fig. 42. IR and ‘H-NMR spectra of 187. (Reproduced by y;6mission of Spriger-Verlag.


Copyright 1991. Reprinted from .)
FURANS IN POLYMERCHEMISTRY 1311

-
HEAT FLOW (mW) ’
EXO 0
t
10 II

-10

200 300 0

TEMPERATURE (“C)

Fig. 43. DSC trace of 190. (Reproduced by permission of Spriger-Verlag. Copyright 1991.
Reprinted from 226.)

5.3.5. Furanic isocyanates

It is well known that certain anionic initiators can induce the polymerisation of isocyanates
through the opening of the C=N double bond to give N-substituted nylons-l. 225These poly-
merisations are often accompanied by trimerisation reactions forming cyclic structures. The
behaviour ‘of two furanic isocyanates, namely 28a and 29a, has been investigated in this
context of Ipolymerisationkyclisation by using a variety of anionic systems.226 Sodium cya-
nide, the ‘“classical” initiator with aliphatic isocyanates, failed to show any activity with
these furanic homologues, probably because the CN- anion was complexed or reacted with
the heterocycle. Both n-butyl lithium and sodium naphthalene were found to operate ade-
quately and gave high monomer conversions at low temperature in various solvents. The
polarity of the medium was a decisive factor for discriminating between the formation of
cyclic trimlzrs 187 and 188 and the corresponding linear polymers 189 and 190.226

(187) (188)
1312 A. GANDINI and M. N. BELGACEM

\ ‘N k \

(189)
n 3-
1’

(190)
n

In fact, solvents with low polarity (e.g. toluene) favoured polymerisation, whereas in strongly
polar media, trimerisation dominated. Fig. 42 shows typical spectra of one of these trimers.
Polymer 190, derived from 2-fury1 isocyanate, had a low DP, but displayed a remarkable
thermal stability. A small fraction of these products was insoluble and highly crystalline with
a melting temperature at 370°C just preceding depolymerisation, as shown in Fig. 43. The
problem here, however, was the difficulty related to the storage of the monomer because of its
already-mentioned instability in the presence of traces of all sorts of atmospheric impurity.
Polymer 189 had a crystalline structure and melted with depolymerisation around 185°C. The
corresponding monomer 29a is quite easy to handle and to store, but the presence of the
methylene moiety between the polymerisable function and the heterocycle clearly weakened
the polymer thermal stability, as already observed and discussed in previous sections, with
furanic polyamides and polyurethanes. These polymerisations should be studied further in
order to optimise the conditions in terms of yields and DP values because they were found to
proceed smoothly and without detectable side reactions.226 In other words, structures 189 and
190 reflected faithfully the regular polymers obtained.

5.3.6. 2-Furjbylidene ketones

The classical products of the condensation of furfural with acetone, 22a and 22h, have
already been examined in the realm of acid-catalysed resinification reactions (see Section
5.2.6). Their behaviour in basic media or with nucleophilic catalysts is quite different and had
only been sporadically examined up until a decade ago. One must distinguish between two
different approaches to this domain, namely the use of aqueous a1kali**‘~***or the initiation
with strong Lewis bases in dry organic solvents.24’229
With the former type of system, applied to 22a, yellow-brown oligomers with molecular
weights ranging from 1000 and 2000 were obtained in 50 to 90% yields, depending on the
specific conditions. **’ These products were soluble in common polar solvents but were only
analysed by infra-red spectroscopy, **‘,**’except for a brief mention of their UV absorption
which occurred roughly at the same -wavelength as for the monomer, albeit with a lower
extinction coefficient.228 The basic qualitative feature distinguishing the monomer from the
oligomers was, however, related to the appearance of a new carbonyl band at 1705 cm-‘. The
polymerisation mechanisms,227~228 put forward on the basis of this scanty evidence, must
therefore be taken with circumspection and would need more work, particularly in the realm
of structural analysis, to be confirmed or modified.
The first study of non-aqueous systems229 was carried out using n-butyllithium in toluene
at room temperature. The authors only worked with 22a, and did not specify its concentration,
or that of the initiator, in the polymerisation experiments. A single yield of 70% was
FURANS IN POLYMER CHEMISTRY 1313

announced for a run lasting about 4 days, but again the specific conditions were not given.
The products were characterised by various spectroscopic techniques, but not in terms of
molecular weight, and it was mentioned that they were poorly soluble in organic solvents.
Three alternative units were envisaged for the polymer structure, arising respectively from a
standard C=C propagation, the opening of the carbonyl group and a Michael-type rearrange-
ment. This investigation had the merit of opening a new area, but lacked the thoroughness
required to establish firm conclusions.
The same problem was taken up a few years later, in our laboratory, with a more compre-
hensive study of the mechanisms and structures involved.24 The use of a variety of model
monomers helped considerably in shedding light on this complex system. A very wide range
of conditions, relative to initiators, solvents, concentrations, temperatures and reaction times,
was scanned in order to obtain a serious insight into the behaviour of the anionic polymer-
isation of 22a in strictly controlled systems in terms of purity of reagents, dryness of media
and uncontaminating atmosphere. The oligomers obtained in these numerous experiments
were soluble in common polar solvents and had molecular weights ranging from 500 to 2000,
depending on the specific polymerisation conditions. A careful spectroscopic analysis of their
structure, based on FTIR, ‘H-NMR and UV data, pointed to different features according to
the type of initiator used. With n-BuLi, the predominant propagation mechanism involved an
isomerisation step with the migration of a proton from the methyl group of the monomer unit
to the ethylenic carbon attached to the carbonyl bond, i.e.

+ Nu - q

etc.

Thus, the “Iphantom” polymer structure in this case was the polyfuryl ketone 191. Apart from
the strong spectroscopic evidence, a clear confirmation of this mechanism was obtained by a
series of experiments with other furfurylidene ketones in which the methyl group was
replaced by the trifluoro (22b), the t-butyl (22d) and the phenyl (22e) counterparts, respec-
tively. 24The absence of a mobile hydrogen atom on the saturated carbon attached to the C=O
group in these structures made these compounds totally inert to anionic polymerisation.
Polymers of 21a prepared above - 30°C with tBuOK or sodium naphthalene gave more
1314 A. GANDINI and M. N. BELGACEM

complex structures which contained units as in 191, but also 192 and, to a very minor degree,
193 and 194.

(193) 1194)
Thus, in this case, the opening of the carbonyl band predominated over the isomerisation
propagation through the C=C moiety. This change in mechanism, induced mainly by tBuOK,
was also corroborated by changing the monomer structure. In fact, 2-furylacrolein 22f poly-
merised with both nBuLi and tBuOK, albeit much more readily with the latter, to give
oligomers made up essentially of unit 195.

60 -

50 -

3000 2500 2000 1500 1000 500

Wavenumbers (cm-‘)

Fig. 44. ITIR spectrum of poly(23a) prepared with tBuOK. (Reproduced by permission of
Elsevier Science Ltd. Copyright 19%. Reprinted from 23’.)
FURANS IN POLYMER CHEMISTRY 1315

In conclusion, 22a responds to nucleophilic attack by two alternative modes of propagation,


depending on the counter ion: the C=C opening with proton shift to give a saturated poly-
ketone, or the C=O opening to give a polyacetal bearing lateral furanalkenyl moieties. The
modest values of DP obtained under all conditions, tested indicate that transfer reactions were
always important in these systems.
The polymerisation of 22a in a basic medium has also been conducted in the presence of
formaldehyde, 230but these systems have not been studied in any detail.

5.3.7. 2-Fwyloxirane

This monomer (23a) is one of the youngest representatives of the furanic family. As
mentioned in Section 3.2.8, its one-pot synthesis from furfural was recently examined and
optimised by using a phase-transfer system. 25*26It is indeed thanks to this qualitative con-
tribution, which makes 23a and its homologues readily available, that several studies have

H3,4

8 i 6 ; 1 i
mm
Fig. 45. ‘H-NMR spectrum of poly(23a) prepared with tBuOK. (Reproduced by permis-
sion of Elsevier Science Ltd. Copyright 1996. Reprinted from 23’.)
1316 A. GANDINI and M. N. BELGACEM

been conducted on the corresponding polymerisations. We have already mentioned briefly its
behaviour and problems related to both free-radical (see Section 5.1 S) and cationic (Section
5.2.7) initiation, 16* but it is with anionic systems that the most interesting results were
obtained.
After a short investigation of the general features encountered in the anionic polymerisa-
tion of 23a with various initiators and on the structural characterisation of its polymers, ‘Q a
more thorough study was carried out on both homo- and copolymerisation systems.231 Bulk
reactions with NaOH or KOH were not particularly satisfactory, whereas tBuOK gave an
interesting situation leading to complete conversion within 3 days at room temperature.
Although the polymerisation did not display an entirely living character, probably because
of transfer reactions, the DP of the polymers increased with increasing monomer-to-initiator
ratio, up to a value of about 100. These products gave NMR and FTIR spectra (see Figs 44
and 45) compatible with a propagation mechanism based on the opening of the oxirane ring,
both through the CY-and the /3-positions, giving units 196 and 197:

a -opening 0
R’ O-M+

0’
s-

(196)

R.
0
p -opening

(197)

The unambiguous establishment of this double mechanism was obtained from alcoholysis
studies on 23a and 23c followed by the spectroscopic analysis of the ensuing addition
products. 232The corresponding macromolecular structures were therefore irregular in terms
of both monomer enchainment and stereospecificity.
When aluminium isopropoxide was used as initiator polymerisations were particularly
active and, more interestingly, gave polymers with a more regular structure. In fact, ar-open-
ing dominated and an appreciable tacticity was also found from the 13C-NMR spectra, as in
the example shown in Fig. 46. Therefore these materials had the specijc structure 198.
.

k-
/
0'

0'
n
(198)
FURANS IN POLYMER CHEMISTRY 1317

C3
(4
I

(b)

C4

%
c,

150 110 70

Ppm
Fig. 46. 13C-NMR spectrum of 198prepared with Al(iPrO)3: (a) as such and (b) in INEPT
mode. (Reproduced by permission of Elsevier Science Ltd. Copyright 1996. Reprinted
from 231.)
1318 A. GANDINI and M. N. BELGACEM

In solution, 23a tended to polymerise more sluggishly when solvents of low polarity were
used. With DMSO and tBuOK at room temperature polymerisations were fast, but reached a
limiting yield, which was higher the higher the initiator concentration. Fig. 47 shows this
behaviour. A kinetic treatment of this system based on a fast initiation step and a termination
involving the active species and a monomer molecule or a monomer unit in the polymer, gave
a satisfactory agreement with the experimental data and provided values of the relevant rate
constants.231
All the polymers obtained in this study showed two major structural features, apart from
the sequential and configurational aspects already discussed, namely:

1. they were in fact telechelic macrodiols (viz. structure 198 with an OH group at each end)
since the acidic neutralisation introduced a terminal OH group in the place of the active
end and produced moreover the hydrolysis of the tBu-0 linkage (in the case of systems
initiated by tBuOK); and
2. they bore the furan ring as a pendant unit without any detectable alteration of its structure,
i.e. no side reactions were involved because of its presence.

The former point is interesting because these macrodiols can be used as a base for chain-
extension reactions, e.g. in the synthesis of polyurethanes. The latter point is particularly
relevant in view of the various problems evoked earlier in the context of both radical and
cationic systems involving furanic monomers.
The polyethers 198 did not crystallise and had glass-transition temperatures ranging from
- 30 to around room temperature as a function of their DP. Their thermal stability extended
up to 300°C.

80

60

0 [Ilo = 0.1220 M
. [Ilo = 0.0797M
m [Ilo = 0.0528 M

0 20 40 60 80 100
Time (mn)

Fig. 47. Kinetics of polymerisation of 23a in DMSO (1.65 M) at room temperature using
tBuOK as inintiator at various concentrations. (Reproduced by YErmission of Elsevier
Science Ltd. Copyright 1996. Reprinted from .)
FURANS IN POLYMER CHEMISTRY 1319

The reactivity of 23a in copolymerisation reactions only allowed the synthesis of copoly-
mers with ethyl oxirane (although even then the furanic comonomer was incorporated in a
predominant fashion), because propylene oxide was too unreactive compared with the 2-fury1
homologue. 23’
The most original feature of 23a was, however, the discovery that it could be oligomerised
by traces of water in the absence of any catalyst.232 The real initiator in these peculiar systems
is in fact the diol80, which is slowly formed by the hydrolysis of 23a. This was confirmed by
the systematic investigation of the possibility of promoting the polymerisation of 23a with
OH-bearing compounds such as alcohols, diols, macrodiols and OH-containing polymers like
cellulose.232 Therefore, the real novelty in this context is the fact that very weakly nucleo-
philic substances like alcohols can activate 23a without the intervention of an added Bronsted
or Lewis base. In the specific instance of water-based initiation, the overall mechanism can
thus be represented as follows:

I \
0
0,
H2O OH 0
- HO b

HO’ + (n - 2) I \
0
T7
0

HO

Careful tlests with different surfaces and specific quenching additives proved beyond doubt
that no external cause could have been an alternative source of initiation. In other words, the
intervention of all possible adventitious presence of traces of acid, base or free-radical species
was ruled out and this left only the direct interaction of OH groups (aliphatic and, even more
strongly, pbenolic) with the oxirane ring of 23a as the origin of the initiation and propagation
steps. 232 The growth of these oligomers, which had DP values of the order of ten, was
therefore the result of a series of additions between an OH-terminated species and a monomer
molecule.
1320 A. GANDINI and M. N. BELGACEM

The obvious implications of this direct addition capability are that various macromolecular
syntheses can be carried out, e.g. the preparation of block copolymers from a macrodiol and
of grafted structures from polymers bearing side OH groups like poly(viny1 alcohol). Un-
fortunately, as already mentioned, the length of the blocks or grafts in these reactions tends to
be limited to about 10 units of 23a, because a specific transfer reaction takes place26 and all
attempts to minimise it have failed up to now.

5.4. Stereospecific polymerisath


The subject related to the possible use of furanic monomers in systems capable of gen-
erating stereoregular polymers is practically virgin. To our knowledge, only one brief in-
vestigation was devoted to this topic.233 It concerned 2-alkenylfurans 21a, 21c and 21d
introduced into typical suspensions of different Ziegler-Natta catalysts in toluene or heptane
at 30 to 50°C. Results with “strong” combinations like Et,Al + TiC14 were disappointing
because the yields did not exceed 12% and the polymers were only partly soluble, suggesting
the occurrence of crosslinking reactions. No stereoregularity was observed in these materials.
With “milder” catalytic combinations, like tBuOAlEt2 + VOC13, up to 30% of polymer was
isolated. The products arising from 21a were again partially crosslinked, whereas with 21c,
linear polymers were obtained with molecular weights of up to 10 000 (viscosimetry) for the
specific catalytic system just mentioned. Spectroscopic analyses showed that these polymers
had the regular structure 199.

21d behaved like 21c and gave the expected structure 170.
All soluble polymers were carefully inspected by X-ray diffractometry and optical micro-
scopy under slow cooling from the melt. No sign of crystallinity was ever detected, suggest-
ing that the search for other stereospecific catalysts is necessary in order to avoid network
formation with 21a and induce stereoregularity with the other monomers. The origin of the
branching reactions responsible for the partial gelling of systems containing 21a is not clear.
It cannot be ascribed to the intervention of the C3, C4 or C5 sites of the furan heterocycle
because 21c and 21d did not suffer from that side event. This conclusion was corroborated by
the fact that 2-methylfuran was not attacked by the catalysts used in this work and was
recovered quantitatively at the end of the blank runs.233 The possible role of the hydrogen
atom on the a-carbon in 21a seems interesting, but was not proved by direct or indirect
evidence.
Therefore, there is still no report of the synthesis of stereoregular furanic polymers, except
for the partial tacticity of the poly(23a) prepared with aluminium isopropoxide (see
Section 5.3.7).
F’URANS IN POLYMER CHEMISTRY 1321

5.5. Conclusions on polyadditions


Much has been done in the field of polyaddition concerning furanic monomers, mostly
derived from furfural, since our previous review. ’ One obvious deficiency in this general area
is the lack. of investigations aimed at preparing stereoregular polymers. Nevertheless, this
increase in knowledge remains globally impressive because it is not only quantitative, but has
often to do with a better understanding of mechanisms and structures, and has therefore
important implications towards properties and possible applications of the corresponding
materials. Some examples of the combination of mechanistic understanding and practical
implications is given in the next section.

5.6. Polymer modifications based on furan chemistry

This selction only describes systems in which novel structures are prepared by using
specific properties of the furan heterocycle. Special sections will be devoted to two major
aspects related to the preparation offuranic polymers which are then modified according to
either photoinduced transformations (see Section 7) or the application of the Diels-Alder
reaction (see Section 8).
It has been repeatedly emphasised that electrophilic media favour substitution reactions at
the C5 position of furan rings. In fact, most cationic polymerisations of furanic monomers, in
which that site is unsubstituted, are accompanied by grafting reactions, albeit with very
variable relative importance (see Section 5.3). If, on the one hand, this side mechanism must
be considered as an unwanted event, because it introduces structural irregularities and even
partial gelation, on the other hand, there are ways of exploiting it advantageously, as indicated
by the foll.owing examples.
The cationic polymerisation of nucleophilic alkenyl monomers is often marred by the fact
that monomer transfer is kinetically important and therefore low DP values are obtained
unless the reaction temperature is drastically reduced. The addition of a simple furan deri-
vative like 2-methylfuran to a cationic system can change this state of affairs because the
reactivity of the C5 position introduces a dominant transfer consisting in the electrophilic
substitution at that site, thus reducing or even practically eliminating monomer transfer. This
concept was applied successfully to the polymerisation of isobutene in the presence of
10.234r235 13y adjusting the reaction conditions in order to optimise the alkylation, oligoiso-
butenes with DP varying from 1 to 100, bearing a terminal furan ring (200) and with an
average functionality of unity, were synthesised and characterised. Thus:

&
+,A- +
(ZOO)

The concept of imposed transfer through the addition of a powerful agent was not new
when this study was conducted, since a similar approach had been applied to the cationic
polymeris,ation of isobutene and styrene in the presence of 2,6-disubstituted phenols.236
However, these systems suffered from two major flaws, namely, the strong retarding role
of the phenolic structure and an efficiency much lower than unity.
1322 A. GANDINI and M. N. BELGACEM

The monotelechelic materials ending with a furan moiety can of course be modified, e.g. by
the Diels-Alder reaction on the furan ring, or by its oxidation with a strong reagent capable of
converting the heterocycle into a functional group. The use of ruthenium tetroxide proved to
be very profitable in this context and turned the furan moiety into a COOH group,234 giving
hydrophobic oligomers with the very polar end group 201:

& + 3.5 RuO, -

+ AcOH + 2c0, + 3.5 RuO,


OH

(201)

These materials are obviously open to further modification or to specific applications when
used as such.
This idea was also applied later, on the one hand, to other monomers, e.g. styrene deriva-
tives, indene, vinyl ethers, and, on the other hand, to a number of other 2-substituted furans
carrying a reactive function, e.g. 51 and 29a, in order to prepare a whole series of interesting
reactive monofunctional oligomers. 235
A logical extension of this trick consisted in using difuran compounds possessing two
equivalent unsubstituted C5 sites, e.g. 12 and its homologues, as specific additives capable uf
inducing a cationic block copolymerisation.235S237 The principle of this convenient one-pot
synthesis is to start with a first monomer A (see scheme below) which will give oligomers like
200, but with a difuranic ending, i.e. with a free C5 position. After A has been consumed, a
second monomer B is added to the system and its oligomeric active species will alkylate those
free sites, thus generating block copolymers 202:

=PolyA+ + DRQ _.
w+nwPol yAaRG

~PolyB+ +
wvuePoly A@($RG -

uur*Pol y AGRa PolyB =


(202)
This principle was applied notably to the isobutene/styrene monomer pair. Conditions had
to be optimised to avoid an excessive double coupling of oligoA and an adequate insertion of
F’URANS IN POLYMER CHEMISTRY 1323

oligoB. The best results gave about 75% of block copolymer with respect to both homo-
polymers and the DP of each block could be varied from 5 to 100, mostly by a progressive
lowering of the reaction temperature. It is important to underline that this technique, which
owes its good performance to the peculiar features of furan chemistry, is not comparable to
the fine tuning and wide scope of living polymerisations, which are capable of providing a
range of bl.ock copolymers with near-quantitative yields and narrow DP distributions. How-
ever, its foremost advantage resides in its simplicity of execution compared with the much
more cumbersome requiremens of living systems.
Apart from the poly(styrene-b-isobutene) copolymers, numerous other materials were
prepared a’ccording to the same procedure, the only limitation being that the monomers used
were exclusively those capable of undergoing cationic polymerisation. 235,238
The very principle discussed above for both monofunctional oligomers and block copo-
lymers applies also to grafting reactions. Thus, for example, a homopolymer or a random
copolymer bearing pendant furan rings with free C5 positions can be introduced in a cationic
polymerisation medium and be the source of grafting, through the electrophilic substitution
reaction of the active species of the monomer being polymerised at the C5 sites. This provides
original macromolecular architectures with variable degrees of branching based on grafting-
onto procedures.
In a completely different vein, propylene-CO alternating copolymers were modified by a
specific reaction involving the transformation of some of their C=O groups into furan hetero-
cycles using P205, to give structure 203:239

_n
& 0 ,m
(203)

The above examples shed some light on the potential interest of using furan chemistry for
the preparation of original polymeric structures. As with other areas discussed previously, we
feel that these openings should stimulate further original investigations.

6. CONJUGATED FURANIC OLIGOMERS AND POLYMERS

The volcanic surge of interest for conjugated polymers in the last 15 years finds ample
justification in their multiple domains of application. 24o In fact, because of their electronic
semiconducting or conducting properties, they are used in electronic transport devices, wave
absorption.. antistatic shielding, etc., and because of their optical and liquid-crystal properties,
they find applications in waveguides, non-linear optical devices, electrochromic, electrolu-
minescent .as well as photoreactive materials. This list is not exhaustive, but simply illustrates
many of the fields related to high technology where conjugated polymers bring a substantial
contribution.
Several families of conjugated polymers have been prepared as amply discussed in two
1324 A. GANDINI and M. N. BELGACEM

recent books,240 starting from the father of them all, namely polyacetylene (204). Aromatic
structures followed, including polyaniline (205) and 1,Cpolyphenylene (206). To these
structures, one must now add “polyheteroaromatic” chains like polypyrrole (207) and poly-
thiophene (208). The next generation of this class of materials comprises the combination of
aromatic (or heteroaromatic) rings conjugated with C=C moieties, such as poly( 1,Cpheny-
lene vinylene) (209) and its thiophene homologue 210.

(204)

Jm
(205)

(206)
As% (207)

The main problems with all these types of structure are difficulties with respect to solubility
and passage to a fluid state (i.e. the lack of easy processing) and the ageing (oxidation,
physical microphase transformation, etc.) of the materials which gives rise to loss of their
key properties. Attempts to solve these obstacles include the introduction of internally plas-
ticising groups such as long lateral aliphatic or oligoether chains. Finally, the conducting
properties of the conjugated polymers must be enhanced by doping, i.e. by adding specific
IWRANS IN POLYMER CHEMISTRY 1325

substances capable of inducing electronic defaults along the macromolecular chain through
the formation of solitons, ions, polarons or bipolarons.
This extremely short walk through conjugated polymers does not really do justice to such a
stimulating field of polymer science, nor does it pretend anything other than to show that
furanic heterocycles have their place in this context. The reader is again referred to the
specialised literature for a thorough treatment of the subject.240 The purpose of this section
is therefore limited to describing the rather rich recent literature on furanic conjugated
polymers, even if it is still a modest contribution compared with the huge number of studies
on the aromatic and other heteroaromatic homologues sketched above.
Two types of oligomeric and polymeric structure are relevant here: those in which the rings
are directly interconnected and those in which they are bridged by alkenyl functions. They
will be treated separately in the following sections.

6.1. Poly(furan)s and copoly(furan)s

Although the cationic polymerisation of furan leads to black materials which exhibit a
certain conductivity, as discussed in Section 5.2.1, those systems are not relevant here
because the macromolecular structures do not contain furanic rings as such, but simply some
conjugated aliphatic sequences. The first report dealing with a material that was assumed to
be poly(furan) 24’ concerned an electrochemical synthesis supposedly leading to four inter-
linked units coupled with a perchlorate anion. Room-temperature conductivities of several
tens of S cm-’ were measured for these films. The evidence for the announced structure was
extremely Itenuous and the term “poly(furan)” seemed therefore a little premature for what
should have corresponded to 211. Similar claims in other studies that followed shortly after-
wards, and which included both electrochemical and chemical initiation, were equally shaky
in terms of structural proofs.2423243In any event, oxidation and other degradation problems
were invok:ed to explain the poor quality and performance of these “poly(furan)s”.244-246

A few years passed and a novel claim of the “formation for the first time of poly(furan) as a
well-characterised regular polyconjugated polymer” was published in a paper which to us
was as poor as the previous reports in proving the linear polynuclear structure 211.247 The
polymer in question, prepared by the electroreduction of 2,5-dibromofuran, gave a h,, at
420 nm and, even after doping with AsFs, displayed a low conductivity of 10e3 S cm-‘. Its
supposedly regular structure containing an enchainment of heterocycles was based on a rather
unconvincing i&a-red spectrum. Moreover, its elemental analysis did not correspond to the
expected structure 211, since it gave a much too low carbon content, which clearly suggested
extensive oxidation, a disturbing fact which was not discussed by the authors.247
It must Ibe emphasised that even the synthesis of oZigo(firun)s 211 with three and four
rings, possessing unambiguously the expected structure, was not achieved until the early
1980s,*4s and this required a cumbersome procedure consisting of successive lithiations and
1326 A. GANDINI and M. N. BELGACEM

couplings with cupric chloride. Before then, only various bifuryls had been reported and their
synthesis had also required rather complicated processes. The preparation of tri- and tetra-
furyls24s opened the way to a more rigorous approach to the determination of the basic
physico-chemical properties of poly(furan) through the study of these well-characterised
oligomers. 249 Later, 250 various trimeric structures, including trifuryl, but also numerous
combinations with thiophene and N-methylpyrrole, were used as monomers in polymerisa-
tions induced by NOPF6, in the hope of generating new doped macromolecules bearing PF;
as counterion. However, all these materials, except those derived from trithiophene, gave
very modest conductivities (around 10m5S cm-‘), probably because they had been overox-
idised during the synthesis.
A more thorough investigation of the electrochemical polymerisation of terfuran25’ pro-
vided interesting information concerning possible structural anomalies with respect to the
regular enchainment 211. In fact the poly(furan) films thus obtained were inspected by IR
spectroscopy with different doping conditions and evidence for ring opening and saturated
moieties was clearly established. These irregularities, brought about by hydrolytic reactions,
were responsible for the poor conductivity and other deteriorating properties observed.
Renewed interest in poly(furan)s, obtained from the monomeric heterocycle both chemi-
cally 252in a neutral form and electrochemically in a doped structure with perchlorate counter-
ions, 253brought little progress to the precise knowledge of the structures and stability of the
resulting macromolecules, which did not seem to us to fit the expected regular polymeric
enchainment 211. Theoretical approaches to the configuration and electronic properties of
poly(furan) and its copolymers with homologous heterocycles252*254-256 have not improved to
any substantial degree the understanding of the properties of these materials.
It has become standard practice to introduce self-plasticising side groups in the structure of
rigid-chain polymers, in order to soften them and make them more readily processable. This
applies equally well to main-chain liquid-crystal polymers and to conjugated polyheteroaro-
matic materials. In particular, this approach has been repeatedly and successfully applied to
thiophene. The polymerisation of 3-alkylfurans was reported to yield “soluble and even
fusible” products,257 but this investigation was not taken any further in terms of structural
characterisation in relation to the expected structure 212. Comparative studies of these
materials with their thiophene homologues258%259led to the observation of photoluminescence
in both series and thermochromic effects only with poly(3alkylthiophene)s 213.

(212) (213)

A recent addition to the family of poly(furan)s called upon the synthesis of trimeric
structures 214, based on a central aromatic ring, bearing various para-positioned substituents,
on which are attached two furan heterocycles. The latter units are linked through their C2
sites to the former, in a paru-configuration. The chemical polymerisation of these trimers
FURANS IN POLYMER CHEMISTRY 1327

failed, but their electropolymerisation gave products which exhibited reasonable conductiv-
ities in their doped state with perchlorate anions.260This study was preceded by a theoretical
examination of the conformation and electronic features of the ideal polymer.261 Regrettably,
again no structural examination of the polymers was conducted to prove that indeed the
materials obtained corresponded to the linear macromolecular enchainment 215.

c?14) (215)

R = R’ = H, CH3, OCH3, OC,H,,


R = 0CH3, R’ = OC,H,,

The very fact that these polymers were insoluble, despite the presence of the side chains on
the aromatic moieties, suggests structural anomalies, i.e. crosslinking reactions either during
or after their synthesis. Despite the insolubility of these materials, which precluded NMR
studies, it is a pity that the authors did not take IR spectra to analyse their structural features,
and confined themselves instead to electrochemical characterisations.
In conclusion, the efforts placed on the understanding of the behaviour of furan and 3-
substituted furans in chemical and electrochemical polymerisations and on the umavelling of
the ensuing macromolecular structures, are far less satisfactory than the corresponding work on
other heterocyclic monomers like thiophenes and pyrroles. It is not clear whether this arises from
intrinsic difficulties at mastering the furanic systems (which tend apparently to give more side
reactions) or simply from a less rigorous engagement in terms of quantitative and qualitative
studies. To our mind, only one investigation251 went to enough trouble to approach these
structural problems in depth and indeed showed that the furan ring in 211 is a fragile entity
which can undergo degradation reactions capable of breaking the conjugation sequences.
Obviously, more work is needed in this area before poly(furan) can gain a rightful place in
the realm of conjugated polymers.

6.2. Poly(2,5-furylene vinylene)


As in the case of the polyaromatic and polyheterocyclic structures discussed in the previous
section, the use of furanic monomers to prepare poly(2,5&ylene vinylene) (216) has received a
minor share of attention.262The first attempt to synthesise 216, i.e. the counterpart of the parent
aromatic and thiophenic homologues 209 and 210, dates from about 15 years ago.263

(216)
1328 A. GANDINI and M. N. BELGACEM

The authors of this investigation took 6 as the monomer and treated it with very strong nucleo-
philes under drastic conditions to induce the successive ‘ ‘aldol-crotonic” condensation of the
methyl and aldehyde functions. They obtained a black product which was insoluble in all
solvents. Its elemental analysis differed very considerably from that of the expected structure
216 and’ its conductivity was extremely low (around IO-’ S cm-‘), although no doping was
attempted. Later work on this material’@ showed that doping it with iodine and ferric chloride
did not improve the conductivity above, at best, 1O-3 S cm-‘. Although it was claimed that the
polycondensation had been successful,263 it is clear that the black insoluble powder was a
crosslinked material with little resemblance to the linear polymer 216.
Several years later, three independent groups published almost simultaneously265-267 a novel
approach to 216, based on the large experience already gathered on the synthesis of 209. The
general route used in these studies, usually called the “precursor technique”, is summarised
below:

In these papers, more emphasis was placed on the characterisation of the intermediates rather
than on the final polymer. In fact, only circumstantial evidence was provided to corroborate
structure 216, without any mention of solubility or molecular weights. However, after doping
with various substances, conductivities of several tens of S cm-’ were reported by all authors.
Work in our laboratory, aimed at repeating these syntheses, failed to give soluble polymers,
despite numerous attempts under various conditions. The products were always black lustrous
brittle films possessing an intractable nature. Indeed, already at the stage of the preparation of
the sulfonium precursor 217, partial precipitation occurred and structural irregularities were
detected. This is hardly surprising, considering the sensitivity of the furan ring to the severe
temperature and nucleophilic conditions used in these preparations.
Recently, more favourable conditions were found in our laboratory capable of leading to
216. 2373262*268
It appeared that the use of 6 in a nucleophilic medium, but less aggressive than
FURANS IN POLYMER CHEMISTRY 1329

that used previously,263 allowed the preparation of oligomers with a regular linear structure.
This investigation started with the optimisation of the synthesis of the dimers 218a and 218b
arising respectively from the self-condensation of 6 and the condensation of 6 with 4.

(218a) (218b)

Both compounds were isolated as truns isomers. Figs 48 and 49 show the FTIR and ‘H-NMR
spectra of 2,18a.
The best conditions for the polycondensation of 6 involved the use of tBuOK in an aro-
matic solvent at 80-100°C. It was soon realised that, if the monomer was introduced in its
totality at the beginning of the reaction, the major products were 218a and low-M’ oligomers,
even when :long reaction times were applied. In other words, the polycondensation ceased in
the early stages of the reaction. This observation led to a mechanistic study in which dimer
218a was treated under the usual conditions applied to 6. No tetramer was isolated and indeed
the reagent was recovered quantitatively. This result, coupled with numerous other pieces of

5o.oc l-

S
9
::
._
E
z

5.00
4000 3500 3000 2500 2000 1500 1000 500

Wavenumbers (cm-‘)

Fig. 48. FTIR spectrum of 218a.


1330 A. GANDINI and M. N. BELGACEM

1 ” ” I ” ” I ” ” I ” ‘. I “. 7 I ” ” 1”‘. II 1” I- ‘, I ” ” 1 -
10 8 6 4 2 0

6 (wm)

Fig. 49. ‘H-NMR spectrum of 218a.

evidence,268 showed that this polymerisation system was a curious combination of step and
chain behaviours. In fact, for the growth of the macromolecule, only those reactions involving
the polymer plus the monomer are possible (as in a polyaddition), whereas the chemical
mechanism associated with this growth is typical of a polycondensation with an AB-type
monomer.
The correct way to conduct this polymerisation was therefore to feed the monomer 6 into
the nucleophilic medium at a very small flow rate so that chain growth (polymer plus mono-
mer) could occur in advantageous competition with respect to dimerisation (monomer plus
monomer) by direct self-condensation.268 In this way, regular 216-type polymers 219 were
obtained with a methyl group at one end and an aldehyde function at the other, as shown by
the following overall mechanism:

t-BuOK
H s

o - n Hz0
(2191
FUR4NS IN POLYMER CHEMISTRY

Fig. 50. ‘H-NMR spectrum of an oligomer 219. (Reproduced by grmission of Elsevier


Science Ltd. Copyright 1996. Reprinted from .)

The monomer units were enchained through tram conformations, as shown by the NMR and
FTIR spectra of samples of oligomers and polymers which resembled those of dimer 218a
(see Figs 48 and 49). The UV-vis. spectra showed a progressive bathochromic shift as the DP
increased, particularly for oligomeric structures. Thus, whereas the dimer absorbed at
393 nm, oligomers with DP values of about 5 gave maxima between 430 and 490 nm and
polymers with DP up to about 50 gave a very broad absorption centred at 5 10 nm, as shown in
Fig. 50. Fig. 51 shows the ‘H-NMR spectrum of a typical oligomer.
It was proved by studies on model compounds that the step leading to chain growth
involves exclusively the aldehyde moiety borne by the polymer (or oligomer) and the methyl
group of the monomer. The highest DP values obtained with these systems were of the order
of 50. These products were brown-reddish powders which gave IR and NMR spectra in
excellent agreement with the expected structure. Z* They were soluble in common solvents
like methylene chloride and did not lose their regular linear structure for years when stored at
atmospheric conditions. After doping with iodine, their conductivity reached values around
0.1 S cm-’ and was quite strongly activated by temperature.268 The photochemical behaviour
of these structures will be discussed in Section 7.
Despite their solubility, the poly(2,5-furylene vinylene)s prepared according to this simple
one-pot procedure were too brittle to allow the casting of films or the spinning of fibres. A
study into the improvement of their processability was undertaken using both physical and
1332 A. GANDINI and M. N. BELGACEM

300 370 440 510 580 650

A (nm)
Fig. 51. UV-vis. spectrum of polymer 219. (Reproduced by pzsmission of Elsevier
Science Ltd. Copyright 1996. Reprinted from .)

chemical means.269 Plasticisation with various classical additives and with oligomers like
PEO failed to give interesting results, mostly because of microphase separation. It was
therefore thought that internal chain softening by block copolymerisation could be an alter-
native solution. This was achieved by taking advantage of the aldehyde end groups on these
chains. 269 After a study on model compounds, condensation reactions were carried out
between the polymers and commercial Jeffamines, i.e. PEO oligomers terminated by one
or two primary amino groups. The result was a series of di- (220) and triblock (221)
copolymers with variable lengths of both the conjugated and the polyether segments.
FURANS IN POLYMER CHEMISTRY 1333

These materials were indeed ‘ ‘plasticised’ ‘, but the operation was excessive and gave
viscous liquids at room temperature. 269 Adjustments are needed to obtain copolymers with
better mechanical properties. The additional interest of these structures is that they contain in
the same macromolecule the aptitude of both electronic and ionic conduction.
All the results discussed above concerning the condensation of 6 apply equally well to the
thiophene homologue 5-methyl-2-thiophene carboxaldehyde, and both random and block
copolymers arising from the combined use of these two monomers were also studied.268
A very exhaustive study of the oligomers of 216 obtained by specific Wittig reactions has
been conducted and reported as a thesis dissertation.*” The synthetic procedures and the
complete spectroscopic characterisation of an impressive number of homologues were
described in that work.
Finally, an alternative route to 216 has been proposed.*‘l It involves the oxidation or
electrochemical polymerisation of truns- 1,2-di(2-furyl)ethylene. However, once again, the
products obtained by applying this procedure were totally insoluble and were not charac-
terised adequately enough to justify the claim of a regular structure like 216. Moreover, the
electronic spectra of these crosslinked materials, as well as their very low conductivities
(even after doping), cast serious doubts about the presence of long conjugated sequences in
their chains.

6.3. Polyenes bearing lateral furan rings

The two previous families of conjugated furanic polymers had the common characteristic
of bearing the heterocycle within the main chain. Even less work has been done on structures
in which the chain is a polyene and the furan ring appears as a side group and these studies are
very recent. They include a short description of the polymerisation of the acetylenic monomer
222 which was aimed at preparing polymers with selective gas permeation.*‘* High mole-
cular weights were obtained with the catalytic system WCl&‘h4Sn and the materials were

\
stable up to 270°C.

F3 CJfI \
0
(222)
CF3

In a different vein, furan moieties were attached to poly(dichlorophosphazene) by con-


densation reactions with sodium 3-furfuryl methoxide273 to give polymer 223:
1334 A. GANDINI and M. N. BELGACEM

(2231

The idea behind this polymer modification was to proceed to a second oxidation step capable
of inducing the junction of the furan rings to form two additional polyconjugated chains, as
shown below:

[Ol

(2241

Whereas the replacement of chlorine atoms with furanic moieties occurred smoothly, the
second step, which should have yielded the cyclomatrix structure 224, was less easy to
optimise. In fact, the resulting dark insoluble products had poor characteristics even when
doped with ferric perchlorate, since their conductivity did not exceed lo6 S cm-‘.
Work is in progress to introduce furanic moieties as side groups to polyacetylene by
various routes which include both the metathesis polymerisation of acetylenic-furanic mono-
mers and the reaction of adequately substituted polyacetylene with specific furanic com-
pounds in order to append variable proportions of the heterocycle on the polyacetylenic
chain. In order to investigate the first type of strategy, a series of new monomers possessing
the general structure 225 were syntbesised and polymerised with typical catalysts like MoCls
and WC16.274 The ensuing polymers were extremely sensitive to atmospheric oxidation, but
FURANS IN POLYMER CHEMISTRY 1335

also to side reactions during the actual polymerisation. Therefore, although there is no doubt
that the propagation occurred through the acetylenic moiety, given the IR spectra obtained,
more work is needed to optimise the conditions of these polymerisations.

R=

In the context of the second approach, poly(butyn-3-01) was prepared in order to be used as
precursor for the condensation reaction with, e.g., furoyl chloride:274

6.4. Conclusions
The potential interest displayed by the introduction of furan heterocycles in conjugated
polymeric structures has only been sketched out by these recent studies. With respect to other
heteroaromatic or aromatic moieties, it offers certain advantages and some drawbacks which
must still be minimised. Also, mixed aromatic/heteroaromatic structures in which each ele-
ment plays a specific role deserve further investigations, particularly with a view to producing
electroluminescent devices. Therefore much remains to be done to prepare materials for, e.g.
electronic, liquid crystal, screening, optical (NLO) and optoelectronic applications.
1336 A. GANDINI and M. N. BELGACEM

7. PHOTOPOLYMERISATION AND PHOTOCROSSLINKING

The photochemistry of simple furans275~276and furfuraL2” among other simple deriva-


tives, is well documented as are the mechanisms of their photooxidation. The photofrag-
mentation reactions are outside the scope of this monograph, whereas the photoaddition
reactions can be relevant to applications to polymer science. The purpose of this section is
to give some relevant examples of polymerisations induced by the electronic excitation of
furan derivatives and of the crosslinking mechanisms arising from the same type of species.
We will limit ourselves to molecular interactions, because the idea of generating free
radicals from the photolysis of a furanic derivative is not in itself a specific feature connected
with the properties of the heterocycle. Thus, for example, there is no qualitative difference
between the use of excited benzoin and furoin to promote a free-radical polymerisation and
therefore the use of the latter does not provide any originality in relation to the actual
polymerisation.
The electronic structure of the furan ring translates into a typical 7~a* absorption
around 200 nm in the absence of any further conjugation. The attachment of a C=C or
C=O moiety to the ring moves this maximum typically to cu. 270 nm. An additional increase
in conjugation through similar groups gives rise to a bathochromic shift of another 70 nm or
so, as in 22a. This progression has been analysed in detail for polyalkylene furans204 and
tends to slow down as the structures lose their planarity. However, it is not difficult to
synthesise furanic moieties absorbing strongly in the near ultraviolet region of the electro-
magnetic spectrum, which are therefore potentially interesting for photoactivated polymer
reactions.

7.1. Photopolymerisations with small molecules

It has already been pointed out in Section 4.1 that furfural can be photopolymerised
in the liquid state to give the linear oligomers 44 resulting from a polyaddition reaction
involving excited furfural and ground-state molecules. 42 Little work has been done on other
systems, apart from classical free-radical polymerisations induced by the photolytic
cleavage of a covalent bond in a typical initiator molecule, as with the photopolymerisation
of 21a.
One novel and interesting approach consisted in making use of the Pate&-Btichi
reaction applied to furan and aromaric diketones. Thus, irradiation of a l/l mixture of
furan and various dibenzoylbenzenes 226 gave positive results, although the materials
obtained had low DP values because of the loss of some furan by volatilisation.279 The
situation was improved by preparing the corresponding adducts 227 from the photochemical
reaction:

(226)
FURANS IN POLYMER CHEMISTRY 1337

226 + 2 0/
0
\ hv ) CR3
/
0
00
OArO

(227)
0
\

These non-volatile intermediates were then mixed in solution with an equimolar amount of
226 and irradiated at 350 nm to give the corresponding polymers 228. The ensuing products
had DP values ranging from 5 to 23 and melted between 150 and 204°C. Their structures were
quite regular, but evidence for minor side reactions (which were responsible for some cross-
linking) was obtained from a careful spectroscopic characterisation. These polymers aged
rather badly, showing the development of carbonyl bands upon storage for several months at
room temperature. Their thermal degradation at 250°C yielded furan as the volatile product
and the aromatic diketone as the residue, i.e. the reverse reaction with respect to the photo-
polymerisat:ion. 279The use of 2,5_dimethyfuran, instead of furan, gave more regular macro-
molecules with no crosslinked side products.

m
(228)
A recent study in our laboratory36 showed that the diacrylic acid 38 can be dimerised in the
solid state b:y irradiation with a conventional tungsten bulb since the monomer absorbs with a
maximum alt 350 nm tailing into the visible:
1338 A. GANDINI and M. N. BELGACEM

The dimer 229 decomposed thermally to regenerate 38 quantitatively, as shown above. The
absorption spectrum of 229 was shifted hypsochromically to 306 nm with respect to that of 38
because of the loss of conjugation. Irradiation of 229 as a solid powder suspended in water
with typical UV sources produced its photopolymerisation which, however, was accompa-
nied by side reactions involving the furan ring. Thus the ensuing polymer was partly cross-
linked. Work is in progress to optimise this reaction and make it as viable as with the
well-known homologous aromatic systems.280 The interest of this piece of work lies, of
course, in the fact that the dimerisation can be carried out in a very clean fashion by using
high wavelengths to which the product is totally transparent. Only a drastic change in the
excitation energy can then promote the second phase, i.e. the polymerisation of 229.

7.2. Photocrosslinking from chromophores on polymer molecules


The general idea of building polymeric structures bearing photosensitive side groups
capable of undergoing crosslinking reaction in a molecular fashion has been exploited for
decades, mostly by using aromatic chromophores. Applications for photographic and printing
processes have become standard practice. 28’ Photocycloaddition reactions involving [2 + 21
intermediates arising from the excitation of (mostly) C=C moieties have been known since
the beginning of the century, e.g. the dimerisation of cinnamic acid in the solid state. The
extension of this concept to photocrosslinking through. photoactive lateral subchains, or
indeed through excitable unsaturation within the main chain, begun about 40 years ago,
precisely with polymeric structures 230 based on poly(viny1 alcohol)s (PVA) which were
substituted to different extents by cinnamoyl moieties through simple esterification reac-
tions.281 The irradiation of this chromophore in the near UV and the intersystem crossing
of the resulting singlet excited state to its triplet electronic configuration corresponded to
creating the biradical intermediate 231 which coupled very readily with a ground-state
structure 230 to give the corresponding cyclobutane-type product 232, i.e.

(230) (231) (232)

Many different polymeric materials possessing these properties have been studied in
detail. *” Most of them are derived from aromatic structures conjugated with one or several
C=C unsaturations and carbonyl groups, but pyridinic counterparts have also received much
FURANS IN POLYMER CHEMISTRY 1339

attention because they can be easily quaternised and rendered water-soluble (before cross-
linking), or at least compatible with polar solvents.28’
The extension of these concepts to furanic derivatives was first reported in the early 1960s
in a general study based on 2-(arylvinyl)-N-vinylpyridinium arylsulfonate polymers.282 In the
specific instance of the introduction of furanic moieties, the reaction leading to the photo-
sensitive polymer 233 was:

R=HorCH,
11n

I
/
N+

\
t-
0'
I

(2331
These hydrophilic polymers crosslinked under UV irradiation and their sensitivity was much
higher than that of the aromatic homologues.
The study of the photochemistry of 2-furanacrylic acids 234283,284 opened the way to
photosensitive polymers similar to the poly(viny1 cinnamate)s, 230. Indeed, tram-234 gives
the corresponding asymmetric cyclobutane: 283

0
(234)
1340 A. GANDINI and M. N. BELGACEM

whereas its cis isomer and various 5substituted derivatives give rise to numerous isomeric
dimers, 283as with cinnamic acids, but also to equilibrium cis-tram isomerisation.284 How-
ever, upon prolonged irradiation at wavelengths higher than 290 nm, 234 also produced
oligomers with DP = 30 which were not analysed in any detai1.283 The photochemical
behaviour of the 2,5_bis(furanacrylic acid) 38 has already been discussed in the previous
section.
Shortly after the study on tram-234, the corresponding polymers derived from PVA were
prepared, characterised and irradiated. 285 In this investigation, furanacryloyl chloride was
made to condense with PVA in a cold NaOH aqueous solution:

The resulting poly(viny1 2-furanacrylate)s


i-7
l 0'
-
0

(235)
0

235 had a high degree of substitution and gave


absorption spectra with a maximum around 300 nm, tailing off all the way to 370 nm.
Irradiation of these materials with a high-pressure Hg lamp gave crosslinked products,
whether or not a sensitiser was used. The actual mechanism of photoreticulation was not
investigated and the only evidence obtained related to the progressive decrease of the IR band
at 1640 cm-‘, suggesting the corresponding consumption of C=C moieties. Regrettably, this
work also failed to give clear information as to the experimental techniques used for the
photochemical, i.e. the basic, aspects of the study.
In another investigation on photosensitive furanic derivatives, the large chromophoric
structure 236, absorbing at 340 nm, was synthesised by a typical Wittig procedure from
furfural and 1,4-dimethyl-bis(triphenylphosphonium chloride). 286Its irradiation gave oligo-
mers that were not characterised in detail.

(2361

The Knoevenagel reaction has also been used to introduce furanic side-chromophores into
polymers bearing C=N and ester functions.287 Furfural was the reagent used for these con-
densations and the ensuing relevant units were like 237:
FURANS IN POLYMER CHEMISTRY 1341

0 0
\
JYAO
F
‘0 I

\’
d
0

(237)
Irradiation of this polymer (which absorbed with a maximum at 335 nm), both as a film or in
solution, was carried out with a medium-pressure Hg lamp on pure samples, but also in the
presence of variable amounts of different sensitisers. Photocrosslinking occurred and the rate of
loss of C=C groups was monitored as judged from the change in the UV spectrum of the polymer
under excitation. No study was conducted on the mechanism of these photochemical eventszg7
Furanic chromophores have also been introduced into amorphous polymeric films contain-
ing active non-linear optical (NLO) moieties in order to induce efficient photoreticulation
effects,288 i.e. the immobilisation of the NLO groups. In this study, it was shown that a fur-
anacrylic function attached to a methacrylate backbone responds more efficiently to this photo-
chemical action than the aromatic homologue. With about 15% by weight of chromophores in
the overall polymeric structure, a maximum of 60% of them could be made to cyclodimerise,
a yield which is already satisfactory in terms of stabilisation of the NLO properties.
In recent years, a systematic investigation has been undertaken in our laboratory to prepare
and assess photoactive furanic structures. 289 We have already discussed in Section 4.4 the
photocrosslinking through the main chain arising from the irradiation of the unsaturated
furanic polyesters 62.74 We also spoke briefly about the photosensitivity of structures 218
and 219, but these topics will now be treated in more detail.
Before looking into the crosslinking of polymers bearing different new furanic chromophores,
a systematic approach to the actual photochemical behaviour of the models was conducted.289
The first series of structures investigated contained only furan moieties conjugated with C=C
and/or C=O groups. A typical example of this comprehensive investigation is the photo-
chemistry of 218a, namely the truns-dimer arising from the polycondensation of 6, described
in Section 6.2. Irradiation of its dilute solutions at room temperature with UV wavelengths
higher than 300 nm induced the progressive conversion of the truns form into the cis isomer
218~ until the thermodynamic equilibrium of interconversion, corresponding to about 2/3 tram
and l/3 cis, was reached.

(218a) (218~)
1342 A. GANDINI and M. N. BELGACEM

As the concentration of 218a was raised, some intermolecular photochemistry began to show
up, but remained a modest event compared with the intramolecular configurational reversion.
Irradiation of this molecule in the solid state, whether dispersed in a KBr pellet or as micro-
crystals suspended in water, led to photodimerisation through a very selective pathway, i.e. that
giving exclusively the symmetrical cyclodimer 238: 289

(238)

Fig. 52 shows the electronic spectrum of a mixture of 218a and 238 in which the partial loss of
conjugation associated with the photodimerisation is clearly indicated by the hypsochromic
shift of the peak for 238.
When 218a had been carefully purified and possessed a high degree of crystallinity, the
yield of dimer 238 was close to 100%. This indicated that the molecular orientation in the
regular cell unit was associated with the ideal spacing, in terms of the distance between C=C
unsaturations, which made coupling between excited and ground-state alkenyl groups occur
smoothly. The cyclodimer 238 was thoroughly characterised spectroscopically. Fig. 53 shows
the ‘H-NMR spectrum of this novel compound, with the simple resonance features which led
to the conclusion about the symmetrical disposition of the methyl and the aldehyde substi-
tuents. Heating 238 above 170°C produced the decomposition of the cyclobutane structure
and the regeneration of the starting monomer.289
The reaction of 218a with pentaerythritol provided the model compound 239 which simu-
lated the attachment of this type of furanic structure on an OH-bearing polymer:289
FURANS IN POLYMER CHEMISTRY 1343

(218a)
.J 0.6

395nm

0.48

0.36
8
9
e
5:
2

0.24

220 300
. I .

380
*
Wavelength (nm)
Fig. 52. UV-vis. spectrum of a mixture of 218 and its cyclodimer 238.

The peculiarity of 239 was to contain two identical independent chromophores per molecule.
Its irradiation in the solid state gave 100% of a single product characterised by a very simple
spectroscopic pattern (absence of alkenyl groups, presence of the typical cyclobutane fea-
tures), as shldwn in Figs 54 and 55, and a clear-cut melting at 144°C. It was concluded that 239
photocyclised through both ends to give the symmetrical cyclophane 240.289
0 0

/o--uxX~
I \
0
0

\I 0 0 \O/
I \
0

L++c~
I\ 0 0 I\
Y \O/ 0
0 0
0
\O/
(240)
1344 A. GANDINI and M. N. BELGACEM

6
6

8 6 4 2

6 (wm)
Fig. 53. ‘H-NMR spectrum of the cyclodimer 238.

This surprising result again points to an extremely favourable intermolecular arrangement of


239 in the crystalline state, such that the probability of polymerisation is negligible, compared
with the double cyclisation through both terminal unsaturations. Note that, since each chro-
mophore here is electronically independent of the other, the cyclisation reaction of one of
them, at one end of the 239 molecule, does not change the photosensitivity of the other. This
is different from the situation discussed above for the furandiacrylic acid 38, which loses
conjugation when it dimerises.
Irradiation of poly(2,5furylene vinylene)s 219 in the solid state produced a partial cross-
linking which was attributed to the crystalline regions, i.e. the microstructures allowing
cyclodimerisation reactions between excited and ground-state unsaturations. The furan rings
did not intervene in these photochemical ramifications and crosslinking. The amorphous
glassy portions of these materials were not affected by irradiation because of the unsuitable
average distance between potentially reactive C=C moieties, which impeded the photo-
coupling and thus dissipated the absorbed electronic energy by vibrational deactivation.
When polymers 219 were irradiated in solution, crosslinking occurred more efficiently and
involved the whole sample. Here, the mobility of the polymer chains gave easy access to the
right intermolecular conformation associated with photocyclisation. 289
A variety of other models was synthesised and tested for their photochemical activity.
FURANS IN POLYMER CHEMISTRY 1345

4000 3000 2000 1000 500

Wavenumber (cm-‘)

Fig. 54. FTIR spectrum of the cyclophane 240.

-1
- I
6
I I
4
I I
2

6 (wm)

Fig. 55. ‘H-NMR spectrum of the cyclophane 240.


1346 A. GANDINI and M. N. BELGACEM

These included many furanic structures, but also the corresponding tbiophenic counterparts
and molecules containing both furanic and pyridinic rings, The latter family was investigated
because of its hydrophilic potential.289 This comprehensive approach included the attachment
of the chromophores to short moieties simulating a reactive polymer.
The final phase of this study289 dealt with the “real” systems, namely PVA variously
substituted with most of these chromophores. Classical acetalysation reactions were used to
append different types and different amounts of photosensitive structures onto PVAs of
different DP values and different degrees of hydrolysis, These functionalised polymers gave
very satisfactory photocrosslinking aptitudes, even when the extent of substitution remained
modest, i.e. below 10%. Two classes of materials were prepared: those bearing non-polar
chromophores, like 241, which were only soluble in non-aqueous solvents; and those posses-
sing pyridinic structures, like 242, which could be made water-soluble after quaternisation. 289

(241) (242)

The maximum absorption wavelength could be modulated and brought very close to the
visible in order to enhance the photoreactivity of these polymers and the interest in their
practical applications. 289

7.3. Conclusions

This very recent incursion into the photochemistry of conjugated furanic structures has
opened a new field of research that is hoped will be taken up and pursued by other groups. Of
course, the implications are not limited to photocrosslinking per se, as indicated by the study
related to improving NLO-type polymers. *”

8. FURAN POLYMERS AND THE DIELS-ALDER REACTION

Of all cycloadditions, the Diels-Alder reaction is perhaps the most emblematic represen-
tative. 290Among the different types of system representing the Diels-Alder reaction, those
involving furan derivatives as dienes are undoubtedly the most classical examples. Indeed, it
can be asserted that the reaction between furan and maleic anhydride is the very symbol of the
FURANS IN POLYMER CHEMISTRY 1347

Diels-Alder reaction. This review is not the place to delve into the actual mechanisms
associated ,with the vast array of inter- and intra-molecular Diels-Alder reactions or into
the classical problems of isomer distribution among products. Two recent books cover these
topics coml>rehensively.290*291 Suffice it here to give a brief reminder of their basic features
and peculiarities: the essential step is a [4 + 21 cycloaddition between a dienic reagent and a
dienophile which are mostly alkenyl or alkynyl derivatives. Thus, in its simplest form, a
Diels-Alder reaction can be represented by the cycloaddition of butadiene and ethylene to
give cyclohexene:

\ 0I
/
+ II -
(
As already stressed on numerous occasions, the furan heterocycle possesses a pronounced
dienic character which manifests itself in Diels-Alder reactions with dienophiles like maleic
anhydride 116, maleimides 243, but also compounds like acetylenedicarboxylic esters 244.
0

c1
/O
N-R

0
(243)
I
0
OR
OR

(244)
Whereas the cycloaddition adducts 117 and 245 are end products, the remaining unsaturation
in 246 can react with one further molecule of furan to give the double adduct 247. Thus 244 is
in fact a bisdienophile. Another example of such a double reagent is the general bismaleimide
moiety 248, which can add onto two furanic rings to give first structure 249 and then the
bisadduct 250:
0

0 + 4-R

(245)

(246)
1348 A. GANDINI and M. N. BELGACEM

OR y
OR

(247)

c3
1
/O O\
N-R-N

0
(248)
0
1

(249) (250)

Of course, whereas the two reactions on an acetylenic bond go through an ethylenic inter-
mediate and are therefore connected successive events, with maleimides, each Die&Alder
cycloaddition is identical in nature and does not depend on whether the other has occurred or
not.
As indicated in all the above interactions, the Diels-Alder cycloaddition has the important
peculiarity of being thermally reversible, i.e. the equilibrium is shifted heavily towards the
reagents as the temperature is raised. Thus, it is usually preferable to carry out Diels-Alder
cycloadditions at moderate temperatures, often with Lewis-acid catalysts.2W’29’ The tempera-
ture at which the retro-Diels-Alder reactions occur depends of course on the specific nature
of the reagents, in terms of both electronic and steric effects.
Not all furanic structures undergo the Diels-Alder reaction.290S29’ Furan and alkyl-sub-
stituted furans show a pronounced reactivity, except when the substituents are too bulky.
Furan-carbonyl derivatives, like furfural and furoic acids, do not react under most
conditions.
How can one exploit this facile and reversible cycloaddition in polymer chemistry? There
are several ways to make use of it, namely:

1. Step polymerisations in which the monomers bear complementary moieties, e.g. a difura-
nit compound (A-A) plus a bismaleimide (B-B);
FURANS IN POLYMER CHEMISTRY 1349

The chemical modification of polymers bearing groups which are prone to the Diels-
Alder reaction, e.g. the use of pendant or terminal furan rings as a source of cycloaddition.
The logical extension of this approach is the crosslinking of linear polymers possessing
reactive moieties by multiple interchain reactions with a complementary bisreagent, e.g.
the use of a polymer with pendant furan heterocycles and a bismaleimide;
The polymerisation of Diels-Alder adducts, e.g. the ROMP mechanism applied to 117 or
245.

These general strategies are obviously common to all structures compatible with the
Diels-Alder reaction but, in the present survey, the only situations that will be discussed
are those involving the intervention of furanic reagents.

8.1. Step polymerisations based on successive Diels-Alder cycloadditions


Apart from several patents dealing with the possibility of constructing polymers by the
reaction of difuranic compounds with comonomers bearing two dienophilic moieties, whose
claims are often unverified, interest in this type of synthetic approach has not been wide-
spread. One of the first scientific studies in this field292 described the preparation of a series of
difuranic compounds linked through their respective C2 positions by different groups,
including aromatic, Schiff base, ester, amino and cyclic acetal moieties. These derivatives
were then submitted to the Diels-Alder reaction with maleic anhydride to generate the double
adducts 25N.

(251)

The subsequent reactions fall outside of the specific context of this section, but can be
analysed to show yet another approach. In fact, the bisadducts 251 were then aromatised in
concentrate’d sulfuric acid and the ensuing bisanhydrides were made to condense with aro-
matic diamines or tetramines to give the corresponding polyimides 252 and polybenzimida-
zoles 253.

251 + H,N-Ar-NH2 _= ZpR$JJArw

(2521
1350 A. GANDINI and M. N. BELGACEM

0.

251 + (H,NbAr-(NH,), -+ ~N~Rf-y+&


-4H20
*N N
(253)

As expected, these polymers, which remained soluble in sulfuric acid, had good thermal
stability as indicated by the onset of decomposition which ranged from 385 to 59O”C,
depending on the structure.292
About 10 years ago, the first example of a synthesis based on an A-A + B-B Diels-Alder
polycycloaddition was published.293 The authors wished to prepare siloxane-containing poly-
imides in order to combine the thermal stability arising from the latter groups with the
flexibility brought about by the former. The A-A monomers were bisfuranic derivatives
254 bearing short siloxane bridges, whereas the B-B monomers were bismaleimides 248
with aliphatic, aromatic or dimethylsiloxane spacers.

\./ \./
R= /\/SLO’Sl~
(254)

Results were quite satisfactory, since the molecular weight of the polymers ranged from
15 000 to 50000. These materials were then aromatised without degradation to give heat-
resistant structures. 293
Interestingly, in a later study,294 this same group found that the difuranic derivative 255
was not reactive, and attributed this failure to the electron-withdrawing effect of the silicon
atoms attached directly to the C2 position of the furan ring. The more conventional aromatic
difuanic compound 256 gave polymers with bismaleimides, but with molecular weights
lower than 10000.294

(255) (256)
FURANS IN POLYMER CHEMISTRY 1351

Oligomers derived from furfuryl alcohol (see Section 4.3) with molecular weights
around 700 have been used as multifunctional monomers in reactions with bis-
maleimides.295 The Diels-Alder cycloaddition took place in this case with a number
of furan rings along the oligomeric chains and therefore produced branched, and finally
crosslinked, products. These materials were submitted to an aromatisation procedure with
acetic anhydride. The only structural characterisation was based on IR spectra which
simply showed the presence of bonds attributed to the monomeric species and the structures
resulting fr’om the cycloadditions. Surprisingly, the thermal stability of these polymers was
modest, showing decomposition from below 2OO”C, even with supposedly aromatised
structures.
The reaction of furfuryl disulfide (257) with various bismaleimides has also been inves-
tigated. 296 This system ought to have produced linear polymers with the general structure
258, since both monomers are bifunctional, but this study reported the formation of dark
powders “insoluble in common organic solvents”, without any interpretation of this feature.
Again, the materials obtained, whether as such or after aromatisation, gave poor thermal
performances, attributed here to the lability of the S-S bond.

(257)

_i#g&+~
(258)

In a different vein, ribbon-like structures have been prepared from the difuranic cyclic
compound 259 and the aromatic bisdienophile 260. 297The structure of the resulting polymer,
with a DP around 7, was characterised very thoroughly and its ladder topology 261
confirmed.
1352 A. GANDINI and M. N. BELGACEM

OW
(259) (260)
R = KH&H, or (CH2)+H3

n
(261)

Another study on the A-A + B-B reaction, involving the Die&Alder cycloaddition and
difuranic monomers, again used bismaleimides as complementary monomers.298 With
difuranic structures bearing a bridge between two C3 positions and methyl substituents at
C2 and C5, results were poor in that the DP values did not exceed 6 at best. This was in
contrast with previous claims of DP values close to 100 for similar systems.293 To us, the
limited extent of cycloaddition seems reasonable in view of the steric hindrance associated
with trisubstituted furanic rings. The use of difuranic monomers linked through the C2
positions without any further substitution gave in fact better results in terms of both yields
and DP values, which were often higher than 20. 298 ‘H-NMR spectroscopy confirmed the
expected polymer structures and showed an excess of exe bicyclic units.
This work also pointed to difficulties related to the aromatisation of the products, under-
lining the fact that previous authors had announced very high yields for this reaction, perhaps
without any real proof to back up their claims. Moreover, these materials displayed a pro-
nounced sensitivity to atmospheric degradation in the form of progressive insolubilisation
caused by radical-induced crosslinking. 298
FURANS IN POLYMER CHEMISTRY 1353

A final piece of work in this general study298 concerned the preparation of poly(ethylene
oxide)s (M’, = 1000-2000) bearing one terminal furanic function and its reaction with bis-
maleimide:s. Again, whereas the 2,3,5-trisubstituted end-heterocycle did not show any apti-
tude to condense with a maleimide function, the 2-substituted homologue reacted well.
A study of A-A + B-B systems is being conducted in our laboratory. Before looking into
the actual Diels-Alder polymerisation, we investigated model reactions involving difuranic
derivatives and monodienophiles, on the one hand, and monofuranic compounds and bis-
dienophiles, on the other.299 Thus, for example, bismaleimide 262 was made to react
with 1 and. 10 and the resulting bisadducts fully characterised. Conversely, various di- and
trifuranic derivatives 263 and 264, with R, R’ and R” being H or CH3, were synthesised and
made to react with 116 and 243. When stoichiometric conditions were used, the products
were shown, by NMR and FIIR spectroscopy, to be mixtures of mono- and diadducts,
the latter being the more abundant the lower the steric hindrance on the furan rings
(R, R’, R” = H).299

(262)

,,!,&Q,., Rtt~R,,
R R’ R R’ R R’
(263) (264)

The polymerisation experiments yielded soluble oligomers (DP = 8) possessing the


expected regular structures arising from Diels-Alder cycloadditions and Tss around 60°C.
Heating these materials produced a weight loss of 5 to 10% around 200°C which was
attributed to loss of terminal furanic units by the retro-Diels-Alder reaction. Thereafter,
the polymers were stable up to 420”C.299 These results show that some (minor) chemical
changes took place in the macromolecular structures, either during storage or while heating
during the TGA analysis, which stopped the occurrence of the retro-Diels-Alder depoly-
merisation along the chains. Aromatisation of the bicyclic adducts is obviously one such
possible event. This is in tune with previous observations concerning the slow atmospheric
“degradation’ ’ of this type of polyadduct.293,294,298
Other A-A + B-B systems are also being scrutinised in order to find the most suitable
combination(s) in terms of both reactivity and high DP values.
We are concurrently studying the possibility of synthesising A-B-type monomers, e.g.
structure 265, and investigating its Diels-Alder polymerisation.
1354 A. GANDINI and M. N. BELGACEM

(265)

Despite previous claims, this compound299 has never been isolated in a pure form. It was
synthesised in our laboratory from the addition of maleic anhydride (116) to furfuryl amine
(18) followed by dehydration of the resulting amino acid in Ac20. Crystalline 265 gave very
clean FTIR and NMR spectra. It melted at 92°C (DSC) and gave an exotherm starting at about
170°C suggesting self-condensation.

8.2. The Die&Alder reactions applied to furanic polymers

The possibility of modifying polymers bearing furan rings through the Diels-Alder reac-
tion is in fact a specific application of the more general topic discussed in Section 5.6. This
aspect was purposely left out then in order to group all systems involving the cycloaddition
reactions in the present context. Two distinct topics will be treated here: (1) the modification
with the aim of obtaining different moieties (related to a corresponding change in properties)
and (2) the crosslinking via the Diels-Alder reaction with a view to making this operation
reversible.

8.2.1. Chemical modijications with monodienophiles

We already mentioned in Section 5.6 the possibility of synthesising oligomers and poly-
mers containing a furanic end group through cationic polymerisations in which predominant
transfer is ensured by added furanic derivatives. 234*235
These monofunctional telechelics were
submitted to simple cycloaddition reactions which introduced novel moieties. A typical
example of this transformation involved the reaction of 2-poly(isobutene)-5-methylfuran
200 with maleic anhydride and the subsequent hydrolysis of the resulting adduct. The
materials thus obtained were therefore hydrophobic chains with two hydrophilic COOH
groups at the same chain end, i.e.
FURANS IN POLYMER CHEMISTRY 1355

H20

xJ@Y?% (266)
lo

0
z

The interest of this type of structure is of course in the realm of surface-active compounds
which can also play, in the specific instance of oligomers 266, the role of viscostatic additives
for oils.
Recent studies have enriched this topic. In a first publication3” this group reported the
synthesis of acrylic oligomers containing furanic end groups and the reaction of the latter with
116 to giv’e polymer 267. The idea was that the adducts thus obtained could play the role of
reactive chain ends, e.g. in polyesterification reaction in which the polyacrylic dianhydride
telechelic would be the acidic macromonomer.

0
30 13

A
0

“Z-(-j 3) 0 I

x-
I
0
~ m
0 0

:k n

(267)

The puzzling feature in this study is that normally a furoic moiety does not react in the Diels-
Alder cycloaddition, whereas the authors claimed a normal condensation which was followed
by UV spectroscopy up to 70% yield.300
The second paper in this series dealt with the preparation of new acrylamides bearing a
furan heterocycle and their radical homo- and copolymerisation.30’ Saturated model com-
pounds of the monomers were also prepared. The Diels-Alder reaction between these models
and 244 only gave the monoadduct despite a large excess of the dienophile, probably because
the reaction was carried out at the relatively high temperature of 80°C. Reactions with the
corresponding polymers were found to be much slower, but no mention of crosslinking was
given, despite the fact that 244 is a bisdienophile.
1356 A. GANDINI and M. N BELGACEM

OH CH; OH

-c-
I
l
CO2

CH;

b a
d

; fi ’
5

8 (wm)
4
LLL
; ; ; il

Fig. 56. ‘H-NMR spectrum of the Diels-Alder adduct: telechelic methacrylic oligomers/
116. (Reproduced by permission of Hiithig & Wepf Verlag, Basel. Copyright 1993.
Reprinted from 3oo.)

The third publication in this series concerns the synthesis of free-radical initiators contain-
ing furanic ends and their use to prepare telechelic acrylic or styrenic oligomers with a furanic
ring at each chain end.302 These materials were then treated with 116 and gave the corre-
sponding Diels-Alder adducts, as shown in Fig. 56.
A fourth contribution from the same group303 describes the synthesis of oligoesters from
116 and 1,4:3,6-dianhydrohexitol and their reaction with mono- and trifunctional furfuryla-
mides to give the corresponding branched or crosslinked adducts.
FURANS IN POLYMER CHEMISTRY 1357

8.2.2. Reactions with bisdienophiles


Early reports on the use of the Diels-Alder reaction to promote crosslinking include the
cycloaddition of difuranic compounds with maleimide-substituted polymers3” and the addi-
tion of a large excess of 116 to poly(vinylfuran),305 although the actual mechanism involved
in the latter system is quite obscure since 116 is a monodienophile. More recently, two
partially hydrolysed poly(N-acetylethylenimine)s grafted with maleimide (268) and furanic
moieties (269), respectively, were made to react by mutual intermolecular cycloadditions.306
This multiple Diels-Alder coupling resulted in a highly crosslinked material (270) following
the reaction scheme:

(269)

(270)

Heating this gel in a good solvent such as nitromethane or MeOWwater, was claimed to induce
the retro-Diels-Alder reaction all the way to the formation of the two initial linear
polymers, i.e.
1358 A. GANDINI and M. N. BELGACEM

270 -
80"C,>2h

r 1

The relatively low temperature at which this furan-maleimide bridging adduct was undone
in 100% yield is quite surprising on the basis of the typical temperature at which the
retro-Diels-Alder reactions take place. Also, the authors did not explain how they
separated the two linear polymers 268 and 269 in order to avoid their recombination after
cooling. 306
Some work in our laboratory has also been devoted to this topic, using different
approaches. In a first preliminary investigation, isobutene was copolymerised cationically
with 2-isopropenyl-5-methylfuran 21d and the resulting materials, with 2 to 10% of furanic
units, were treated with various bisdienophiles in solution. 307Gelation took place readily and
was, of course, faster the higher the furan content in the copolymer. Heating these gelled
systems up to 160°C only brought about a partial dissolution of the product, i.e. some
material remained crosslinked. We attributed this phenomenon to changes in the bridging
adduct structure, leading to irreversible linkages with respect to a total yield of retro-
Diels-Alder.
In a later study, 299,3o8various copolymers of furfuryl acrylate (25a) and methacrylate (25b)
with acrylic comonomers and styrene were prepared and fully characterised. Their content of
furanic units was varied between 3 and 50% and their molecular weights were close to 20 000.
All these materials were treated with a threefold excess of bismaleimide 262 in THF at 40°C.
The expected double Diels-Alder cycloaddition reaction, occurring mostly by interchain
bridging, did take place, as shown by the progressive increase in the solution viscosity all
the way to the formation of a gel, i.e.
FURANS IN POLYMER CHEMISTRY 1359

Fig. 57 shows the FTIR spectrum of the isolated gel, after extraction with methylene
chloride, related to a styrene (9O)/furfuryl methacrylate (10) copolymer. The most important
feature of this spectrum is the virtual absence of the peaks arising from the furan heterocycle,
which proves the near-completion of the cycloaddition.299V308 Heating these crosslinked pro-
ducts in a TGA capsule showed a thermal stability which depended on the content of furanic
moieties, i.e. on the crosslink density: the onset of decomposition varied from 250 to 400°C
when passing from 3 to 50% of furanic units in the original copolymers. Work is underway to
investigate the retro-Diels-Alder reaction with these networks.
In a subsequent study,308 many of the furanic polyurethanes described in Section 4.6 were
submitted to the Diels-Alder reaction. Before turning to crosslinking systems, these poly-
mers were treated with monomaleimides and the resulting products examined spectroscopi-
cally and by elemental analysis. Again, the characteristic IR bands of the furan ring
disappeared (or were weakened considerably) and were replaced by peaks arising from the
corresponding bicyclic adduct structure, as shown in Fig. 58. These results were corroborated
by the corresponding ‘H-NMR spectra, as shown in Fig. 59. The ease with which this model
cycloaddition took place depended on the actual position of the furan ring with respect to the
polyurethane main chain. Thus with pendant heterocycles (as with 89) and with rings in the
1360 A. GANDINI and M. N. BELGACEM

18

16

3
!l
0
, I I I I I
4000 3500 3000 2500 2000 1500 1000

Wavenumbers (cm-‘)

Fig. 57. FTIR spectrum of a styrene (90)/furfuryl methacrylate (10) copolymer crosslinked
with 262.

3,4-enchainment (as with SS), reactions reached completion, whereas with rings in a 2,5-
enchainment, yields ranged from 30 (as with 91) to 60% (as with 87), as an inverse function
of the steric hindrance of the groups neighbouring the heterocycle.308
The furanic polyurethanes were then treated with a twofold excess of bismaleimide 262 in
THF or THFLDMAc at 40°C. Gelation took place in all instances, but the time required for
crosslinking varied very considerably, following the same steric criteria discussed above for
the monocycloaddition. Thus, only a few hours were required to attain infinite viscosity with
polymers 88 and 89, whereas it took up to several days with the more hindered furanic
moieties. The course of these reactions was followed by SEC as shown in Fig. 60: the
progressive shift of the polymer peak towards shorter retention times indicates the corre-
sponding increase in molecular weight related to interchain cycloadditions. Moreover,
obviously, the gelled portion is not eluted and this is revealed by the apparent increase in
the amount of 262 with reaction time; what is in fact happening is that the fraction of soluble
polymer decreases more rapidly than the consumption of 262, which was actually present in
excess. The Tss of the crosslinked polymers increased by 50 to 60°C with respect to those of
the starting linear structures.308
A third piece of work that is currently in progress in our laboratory concerns the acetalysa-
tion of poly(viny1 alcohol) with furfural and the reaction of these modified polymers
with dienophiles. Furfural reacts readily with PVA both in aqueous and DMSO solutions
and up to 60% of the OH functions can be acetalysed.309 These polymers were thoroughly
characterised and Fig. 61 shows a typical ‘H-NMR spectrum. Their reaction with 116,
FURANS IN POLYMER CHEMISTRY 1361

4000 3500 3000 2500 2000 1500 1000

Wavenumbers (cm-‘)

Fig. 5 8. FTIR spectrum of the polyurethane 89: (a) before and (b) after reaction with 243
(R = CH3).

243 and 262 is being investigated, together with the corresponding retro-Diels-Alder
behaviour Iof each system.
Finally, another system has been investigated, namely the synthesis of styrenic-furanic
random copolymers possessing the general structure 271: 310

4
r&t
0

0
0
n

(271)
5-
0'
m
1362 A. GANDINI and M. N. BELGACEM

(b)

OJ,
‘0
(a) 8’L/ 9’

Hz0

7’+8’ 4

,1
9’

.J

Fig. 59. ‘H-NMR spectrum of the polyurethane 89: (a) before and (b) after reaction with
243 (R = CH3).
FURANS IN POLYMER CHEMISTRY 1363

b
..

0 10 20 30 40

Retention time (mn)


Fig. 60. SEC of the reaction of polyurethane 94 with 262 at different crosslinking times: (a)
t = 0, (b) t = 1 day and (c) t = 2 days.

These materials were prepared by two alternative routes: (1) the radical copolymerisation of
styrene with the corresponding furanic homologue, prepared by the reaction of 4-cblorometbyl-
styrene (272) and 7, or (2) the radical copolymerisation of styrene with 272 followed by the
reaction of the resulting copolymers with 7 to replace the chlorine atom by the furfuryloxy
moiety. 310Fig. 62 shows the ‘H-NMR spectrum of one of these copolymers 271. The reaction of

/
these materials wirhmonodienophiles showedthe ready availability of the furan ring and thereafter
crosslinking with bismaleimide 262 was studied as well as the retro-Diels-Alder of the ensuing
networks. The latter reaction was conducted in the presence of an excess of 10,
in order to trap the
released dienophile and avoid re-crosslinking upon cooling.

6 0

(272)
Cl
1364 A. GANDINI and M. N. BELGACEM

2 2 2

040 OH

0 \); 13

8 6 4 2

6 (w@

Fig. 61. ‘H-NMR spectrum of a PVA acetalysed by furfural.

8.3. The ROMP applied to furanic Die&Alder adducts

Ring-opening metathesis polymerisations (ROMPS) have acquired great notoriety in


recent years because they are a source of interesting materials and can be conducted in
certain instances with very simple means, e.g. aqueous systems without any particular
protective measure.3” Bicyclic alkenyl structures have thus been studied extensively.31’
The polymerisation of bicyclic compounds derived from the reaction of furan and its
derivatives with 116 and 243 is a recent important addition to the existing large body of
literature on olefinic monomers. The first reports in this field described the polymerisation of
a derivative of adduct 117 and of adduct 246 induced by osmium trichloride or ruthenium
trichloride, to give polymer 273.3’2Y3’3Thus, for the latter system:

x
0 0

‘0 - 0’
0
n

(246) (273)
FURANS IN POLYMER CHEMISTRY 1365

(273) (274)

2 2 2 2
1 1

- T a0 a a a
a a a0 a
b
4
4

‘1
1
i -gy

J
15.59 160.17 59:28 al.18
I t I ” I”, 1 I “I ’ I s 8 I ’ II 8 1, I r * I I 17 8 7 - I, 9 - ) I” 3
9 8 7 6 5 4 3 2 1

6 (PPN

Fig. 62. ‘H-NMR spectrum of a copolymer 271.


1366 A. GANDINI and M. N. BELGACEM

Me
CDCI,
MeO,C CO,Me c+t
\ / Ii=

160 1.50 140 130 120 110 100 90 80 70 60 50

Fig. 63. 13C-NMR spectrum of polymer 273. (Reproduced by permission of the Royal
Society of Chemistry. Copyright 1993. Reprinted from 3’3.)

Fig. 63 shows the 13C-NMR spectrum obtained with the former catalyst.313 Interestingly, this
polymer was dehydrogenated to what was claimed to be the conjugated homologue 274, i.e. a
paramagnetic material absorbing at 460 nm, although little evidence was provided for that
structure. Conversely, the monomer was partially reduced to compound 275 which was then
polymerised to give the tetrahydrofuranic polymer 276. 3’3
In a similar vein, both the hydrolysed (277) and reduced (278) derivatives of adduct 117
were polymerised by ROMP to give the corresponding polymeric structures 279 and 280. 3’4
Again, the structure of these materials was carefully examined.

00
H! 'H
(279)
/
FURANS IN POLYMER CHEMISTRY 1367

rx’
O

cc
10 z’
\
-
P9
n

(278) (280)

A similar study was conducted at the same time on the adduct of 1 with N-methyl-
maleimide and emphasis was placed on the thermal stability of the high-molecular-weight
polymer 281 thus obtained.3’5

(281)

The Diels-Alder reaction of 1 with hexafluorobut-Zyne gives the bicyclic compound 282,
which was activated by ROMP with ruthenium trichloride hydrate and gave polymer 283.
This material had a Tg of 93°Cand decomposed at 260”C.316

(282) (283)

Other 2,3_disubstituted 7-oxanorbomenes and 7-oxanorbornadienes have also been studied


in detail wrth molybdenum-based catalysts. 3’7
More recently, the furan/maleimide adduct has been polymerised by aqueous ROMP.3’8V319
The use of ruthenium trichloride, proposed in earlier studies, was revived and more emphasis
was placed on studying the polymerisation kinetics.318 The interest of this work resides
moreover in the extreme simplicity of the systems, which require water as solvent and readily
available commercial initiators. These investigations were complemented by the synthesis of
a copolymer between 117 and 246.320 The same group also studied the partial depolymerisa-
tion of structures 273, 281 and 283 to oligomers bearing trimethylsilyl end groups.32’
Polymer 281 can also be prepared by treating the corresponding monomeric adduct with
methyltrioxorhenium/AlClEt2 in non-polar solvents.**
1368 A. GANDINI and M. N. BELGACEM

Substituted 7-oxanorbornadienes like 246 have also been polymerised in a more conven-
tional way, i.e. through the opening of the unsubstituted C=C unsaturation:“22.“2’

@g - I$$
00

0
(284)

The use of specific Pd complexes provided fast and complete polymerisations possessing a
living character. Moreover, alternating copolymers with CO were also prepared:

Heating both 284 and 285 induced the retro-Diels-Alder


w0
/O

0
00
-

(285)
n

reaction which released the 3,4-


furandicarboxylate and left polyacetylene from 284 and poly(acetylene-a/t-CO) from
285 323,324

8.4. Conclusions

The very fact that the furan heterocycle is one of the most classical representatives of
dienic reagents for the Diels-Alder cycloaddition makes the exploitation of this reactivity in
the realm of polymer chemistry a fertile field of research. Considering this potential, it is
surprising that studies related to this topic should have started so late and so sluggishly. In
recent years, the situation has evolved towards a more vigorous activity. Two types of system
seem particularly promising, namely, the possibility of achieving reversible crosslinking,
which would be particularly useful for preparing thermoplastic elastomers (e.g. recyclable
tyres), and the new class of materials arising from the application of ROMP to various
bicyclic adducts.
FURANS IN POLYMER CHEMISTRY 1369

9. THE AGEING OF FURANIC POLYMERS

In the previous review on furanic polymers, ’ a section was devoted to the “Stability and
Resinification of Furan Derivatives’ ’ . It was underlined in that context that the nature of the
substituent {directly connected to the heterocycle played a key role in determining the
reactivity of the resulting compound towards atmospheric and storing agents, i.e. oxidation,
photodegradation and acid- or base-catalysed transformations.
In practice, labile hydrogen atoms on a carbon atom attached to the furan ring are the most
likely candidates to initiate these degradation processes. Other reactive moieties include chemi-
cal functions that display a tendency to induce intermolecular condensations. An example of such
a behaviour is found with 2-furfuryl chloride, in which the dienic character of the heterocycle is
reflected in the enhanced polarity of the C-Cl bond and the consequent tendency of electrophilic
substitutions between this moiety and the C5 position of another molecule. This situation leads to
black resinous materials (resembling of course those obtained with 7) from the bulk liquid and
the only way to store this compound is therefore in a dilute solution at low temperature.
When the chemical functions attached to the heterocycle are carbonyl or carboxyl groups,
no major problem of degradation is usually observed and therefore compounds like 2-furoic
acid and its esters can be stored for years without any detectable changes.
These considerations can be applied to polymers bearing furan rings. A few examples will
illustrate the very variable responses of these materials to ageing, i.e. to degradation under
natural or more severe environments. High-molecular-weight poly(Zvinylfuran) 129 has a
pronounced tendency to become insoluble on standing, e.g. after precipitation and drying.
The origin of this crosslinking phenomenon must be found in the ease with which the tertiary
hydrogen atom can be abstracted, i.e. with the relative stabilisation of the resulting free radical:

(286)
The driving force for this reaction can be, as with all typical degradation mechanisms, electronic
excitation by absorption of near-UV daylight, or the presence of another free radical (e.g. from
oxidation reactions) which will incorporate the hydrogen atom through the formation of a
covalent bond. Subsequently, the macroradicals 286 can couple, thus inducing first branching
and ultimately crosslinking:
1370 A. GANDINI and M. N. BELGACEM

This overall mechanism is not new or original in itself, since most vinyl-type polymers with a
stabilising substituent are known to display the same behaviour, e.g. poly(styrene). What differ-
entiates 129 from its non-furanic homologues is the kinetic aspect of the phenomenology: the
fact that the reactions involved occur much more rapidly because of the dienic character of the
heterocycle, which plays an “autocatalytic” role in the process. These degradative events are
also observed in furanic polymers in which the ring is within the macromolecular backbone and
bears methylene moieties. In this case, the fragility of the materials becomes apparent particu-
larly upon heating, i.e. their thermal stability is usually quite poor.
For reasons that were explained in great detail in the appropriate sections, furanic polymers
bearing heterocycles with unsubstituted C5 sites are prone to suffer branching and cross-
linking in acidic media, or simply if stored in an environment which is slightly electrophilic.
Although of course most of these degradation processes are detrimental to the properties
and applications of the corresponding materials and require the use of appropriate stabilising
additives, in some instances the products arising from these mechanisms possess interesting
properties that can be advantageously exploited, as pointed out above in various specific
systems, e.g. the nucleophilic resins from poly(2-vinylfuran).
To our mind, the real problem here is not so much the occurrence of the ageing features,
even when they take place at a high rate, but rather the lack of studies on how to inhibit, or at
least retard them. It seems, in fact, logical to envisage that appropriate stabilising agents, such
as those currently used for conventional polymers, should solve these problems. In other
words, instead of discarding certain furanic polymers on the grounds of their fragility, there
ought to be work done on their stabilisation. Many materials would certainly turn out to be
quite viable after suitable investigations directed at finding adequate additives. The example
of PVC, a very fragile polymer which has nevertheless found so many important applications
thanks to appropriate stabilisation, is a remainder of the validity of the above considerations.
In order to dispel false “common knowledge”, it is important to remind the reader that
together with the “delicate” polymers just described, many other furanic polymers show a
remarkable intrinsic stability and may be used as such without hesitation. Examples of such
materials, as thoroughly described in the preceding sections, are polyesters, polyamides and
polyurethanes in which the heterocycles within the chain are mostly present with carbonyl
groups attached at both its ends, whether at C2/C5 or C3/C4. In these instances, there is no
difference in their ageing compared with that of the best aliphatic and aromatic counterparts.
Thus, the unique advantages brought about by the presence of the furanic heterocycle in these
structures is not perturbed by any detrimental disadvantage and the materials offer as a
consequence new fields of added-value applications.

10. GENERAL CONCLUSIONS


Renewable resources under a variety of forms, i.e. wood, annual plants, but also agricul-
tural and forestry wastes, represent extremely precious raw materials for the implementation
of a novel strategy of polymer synthesis that could replace progressively the processes based
on fossil resources. The present review has focused exclusively on one facet of this strategy,
i.e. that related to furan chemistry applied to polymer science. The exploitation of polysac-
charides, lignins, terpenes, tannins, oils, etc. represent the rest of this vast field of research and
development.
Even within the restricted area of this survey, the number of different approaches and
F’URANS IN POLYMER CHEMISTRY 1371

corresponding materials is already quite astounding. Notwithstanding the unifying aspect of


the specific features related to the furanic heterocycle, there exists a very wide spectrum of
possibilities in terms of chemical operations and consequently of resulting structures and
properties.
Our aim in writing this review was to emphasise: (1) the variety of studies and results
already attained, which show that numerous and unique materials have already been pre-
pared; and (2) the yet unexploited potential of this field, i.e. the fascinating and useful work
that lies ahead. If more laboratories should feel a motivation to become engaged in this
research darmain after reading this survey, the message we wished to convey will have
reached its goal.

REFERENCES
1. Gandini, A., Adv. Polym. Sci., 1977, 25,47.
2. Gandini, A., in Comprehensive Polymer Science, First Supplement, eds. S. L. Aggarwal and S.
Russo. Pegamon Press, Oxford, 1992, p. 527.
3. Gandini, A. and Belgacem, M. N., The Polymeric Materials Encyclopedia, Vol. 11, ed. J. Sal-
amone. CRC Press, New York, 1996, p. 8518.
4. McKillip, W. J., ACS Symp. Ser., 1989, 385,408.
5. Dunlop, A. P. and Peters, F. N., The Furans. Reinhold Publishing Co., New York, 1953.
6. Bird, C. W. and Cheeseman, G. W., in Comprehensive Heterocyclic Chemistry, Vol. 4, eds. A. R.
Katritzky and C. W. Rees. Pergamon Press, Oxford, 1984, Ch. 1, 2 and 3; Dean, F. M. and
Sargent, M., in Comprehensive Heterocyclic Chemistry, Vol. 4, eds. A. R. Katritzky and C. W.
Rees. Pergamon Press, Oxford, 1984, Ch. 10 and 11; Donnelly, D. M. X. and Meegan, M. J., in
Comprehensive Heterocyclic Chemistry, Vol. 4, eds. A. R. Katritzky and C. W. Rees. Pergamon
Press, Oxford, 1984, Ch. 12.
7. McKillip, W. J. and Sherman, E., in Kirk-Othmer Encyclopedia of Chemical Technology,
Vol. 11, ed. M. Grayson. John Wiley & Sons, Inc., New York, 1981, p. 499.
8. Theander, 0. and Nelson, D. A., Adv. Carbohydr. Chem. Biochem., 1988,46,273.
9. Kustex, B. F. M., Starch, 1990, 42, 3 14.
10. Fuchs, A., paper presented at 3rd Seminar of Inulin, International Agricultural Centre, Wagening,
The Netherlands, 1989.
11. Penczek, S. and Kubisa, P., in Comprehensive Polymer Science, Vol. 3, eds. G. C. Eastmond, A.
Ledwith, S. Russo and P. Sigwalt. Pegamon Press, Oxford, 1992, p. 772.
12. Dreyfuss, M. P., Poly(tetrahydrofuran). Gordon and Breach, New York, 1982.
13. Brown, W. H. and Sawatzky, H., Can. J. Chem. 1956, 34, 1147; Yoshiaki, K., Katsumi, H.,
Koichi, H., Masahiro, A., Yoshinori, N. and Junji, F., J. Am. Chem. Sot., 1956, 98, 1147.
14. Timko, J. M., Moore, S. S., Walba, D. M., Hiberty, P. C. and Cram, D. J., J. Am. Chem. Sot.,
1977,99,4207.
15. Pennanen, S. and Nyman, G., Acta Chem. Stand., 1972,26, 1018.
16. Holfinger, M. S., Conner, A. H., Helm, D. R. and Hill, C. J., J. Org. Chem., 1995,60, 1595.
17. Skouta, M., Lesimple, A., Le Bigot, Y. and Delmas, M., Synth. Commun., 1994, 24, 2571.
18. Cawse, J. L., Stanford, J. L. and Still, R. H., Makromol. Chem., 1984, 185,697.
19. Quillerou, J., Belgacem, M. N., Gandini, A., Roux, G. and Rivero, J., Polym. Bull., 1989,21,255.
20. Alvarez, R., Gandini, A. and Martinez, R., Makromol. Chem., 1982, 183, 2399.
21. Arekion, J., Delmas, M. and Gaset, A., Biomass, 1983, 3, 59.
22. Alvarez, R., Gandini, A., Martinez, R., Ortiz, P. J. and Perez, C. S., Revista Cenic, 1974.5, 179;
Chem. Abstr., 1975,82, 111867;Gandini, A. and Martinez, R., Revista Ciencias Quimicas, 1983,
14, 111.
23. Isacescu, D. A. and Avramescu, F., Rev. Roum. Chim., 1978,23, 873.
24. Salon, M. C., Gandini, A. and Cheradame, H., Polym. Bull., 1984, 12,441.
25. Borredon, E., Delmas, M. and Gaset, A., Tetrahedron, 1987, 17, 3945.
1372 A. GANDINI and M. N. BELGACEM

26. Boufi, S., Belgacem, M. N. and Gandini, A., Polym. J., 1997, 29, 479.
27. Gandini, A. and Rieumont, J., Tetrahedron Z&t.., 1976, 25, 2101.
28. Rieumont, J. and Espinosa, R., Revistu Ciencias Quimicas, 1983, 14, 285.
29. Labidi, A., Salon, M. C., Gandini, A. and Cheradame, H., Polym. Bulk, 1985, 14, 271.
30. (a) Watanabe, W. H. and Conlon, L. E., J. Am. Chem. Sot., 1957,79, 2828. (b) Brunache, M. H.,
Doctorate Thesis, National Polytechnic Institute, Grenoble, France, 1984.
31. Blomquist, A. T. and Stevenson, H. B., J. Am. Chem. Sot., 1934,56, 146; Stevenson, H. B. and
Johnson, J. R., J. Am. Chem. Sot., 1936,59, 2525; Singleton, H. M. and Edwards, W. R., J. Am.
Chem. Sot., 1938, 60, 540; Garcia, T. and Macias, A., Revistu Ciencius Quimicas, 1981, 12, 2;
Chem. Abstr., 1982,98,3498 1; Masson, M. A., Bouchy, A., Dormoy, J. R. and Roussy, G., J. Mol.
Struct., 198 1, 74, 11.
32. Garber, J. D., US Patent No. 3 049 552, 1962; Quillerou, J., Doctorate thesis, National Polytechnic
Institute, Grenoble, France, 1991.
33. Nielek, S. and Lesiak, T., in Topic in Furan Chemistry, ed. J. Kovak. 4th Proc. Symp. Furan
Chem., Slovak Teck. Univ., Bratislava, Czechslovakia, 1983, p. 183; Chem. Abstr., 1983, 101,
230248.
34. Hirai, H., J. Mucromol. Sci., Chem., Part A, 1984, 21, 1165.
35. Mouloungui, Z., Delmas, M. and Gaset, A., Synth. Commun., 1984, 14,701.
36. Viallet, A. and Gandini, A., J. Photochem., 1990, A54, 129.
37. Roque-Malherbe, R. and De Onate-Martinez, J., J. Muter. Sci. L&t., 1993, 12, 1037.
38. Galego, N. and Gandini, A., Revista Cenic, 1975, 6, 163; Galego, N. and Gandini, A., Revistu
Cenic, 1984, 15, 143.
39. Sanchez, R., Hernandez, C., Jalsovszky, G. and Czira, G., Eur. Polym. J., 1994, 30, 37.
40. Sanchez, R., Hernandez, C. and Keresztuty, G., Eur. Polym. J., 1994, 30,43.
41. Sanchez, R. and Hernandez, C., Eur. Polym. J., 1994,30,5 1.
42. Galego, N. and Gandini, A., J. Polym. Sci., Polym. Chem. Edn, 1977, 15, 1027.
43. Rodriguez, V. J. and Gandini, A., Revista Cenic, 1974, 5, 29.
44. Rodriguez, V. J. and Gandini, A., Revistu Cenic, 1975, 6, 155.
45. Burma, S. P. N. S., Lenka, X and Nayak, P. L., J. Appl. Polym. Sci., 1992, 44, 1311.
46. Prusty, J., Lenka, S. and Nayak, P. L., Mucromol. Rep., 1993, A30, 27.
47. Swain, S. K., Shadoo, S., Mohapatra, D. K., Mishra, B. K., Lenka, S. and Nayak, P. L., J. Appl.
Polym. Sci., 1994, 54, 14 13.
48. Gandini, A., in Encyclopedia of Polymer Science and Engineering, Vol. 7, eds. H. F. Mark, N. M.
Bikales, C. G. Overberger and G. Menges. Wiley, New York, 1988, p. 454.
49. Schmitt, C. R., Polym. Plast. Technol. Eng., 1974, 3, 121.
50. Choura, M., Belgacem, M. N. and Gandini, A., Macromolecules, 1996, 29, 3839.
51. (a) Wewerka, E. M., J. Polym. Sci., 1971, A19, 2703; (b) Barr, J. B. and Wallon, S. B., J. Appl.
Polym. Sci., 1971, 15, 1079.
52. Fawcett, A. H. and Dadamba, W., Mukromol. Chem., 1982, 183,2799.
53. Gonzalez, R., Martinez, R. and Ortiz, P., Makromol. Chem., 1992,193,l; Gonzalez, R., Martinez,
R. and Ortiz, P., Makromol. Chem., Rapid Commun., 1992, 13, 5 17.
54. Chuang, I.-S. and Maciel, G. E., Ann. Rep. NMR Spectrosc., 1994, 29, 169.
55. Glowinski, S. and Pajak, Z., Acta Phys., Polym., 1978, A54,411.
56. Nguyen, C. and Zavarin, E., J. Wood Sci. Technol., 1986, 6, 15.
57. Nguyen, C. and Zavarin, E., J. Appl. Polym. Sci., 1992, 46, 1103.
58. Fawcett, A. H., Surgenor, A. E., Bashir, Z., Litchfield, R. E. and Todd, N. G., Polymer, 1992,33,
5282.
59. Ma, C. C. M., Goang, D. Y., Han, J. L. and Hsieh, K. H., Angew. Mukromol. Chem., 1994,214,39.
60. Ma, C. C. M., Goang, D. Y., Han, J. L. and Hsieh, K. H., Angew. Makromol. Chem., 1994,218,
41.
61. Spange, S., Heublein, B. and Schramm, A., Makromol. Chem., 1992,193, 1; Spange, S., Schtttz,
H. and Martinez, R., Makromol. Chem., 1993, 194, 1537; Spange, S., Winkelmann, H. and
Martinez, R., Angew. Mukromol. Chem., 1993,208, 125.
62. Mariwala, R. K. and Foley, H. C., Ind. Eng. Chem. Res., 1994, 33, 607; Lafyatis, D. S.. Tung, J.
and Foley, H. C., Ind. Eng. Chem. Res., 1991, 30, 865.
FURANS IN POLYMER CHEMISTRY 1373

63. Yamada, S., Sato, H. and Ishii, T., J. Compos. Mater., 1973, 2, 253.
64. Manocha, L., Yasuda, E., Tanabe, Y. and Kimura, S., Carbon, 1988,26, 333.
65. Imito, H., Omaru, A., Azuma, H. and Nishi, Y., Proc. Electrochem. Sot., 1993,93,9.
66. Glowinski, S. and Pajak, Z., Carbon, 1978, A54,411.
67. Shindo, A. and Izumino, K., Carbon, 1994, 32, 1233.
68. Moore, J. A. and Kelly, J. E., Macromolecules, 1978, 11,568.
69. Moore, J. A. and Kelly, J. E., J. Polym. Sci., Polym. Chem. Edn, 1978, 16,2407.
70. Moore, J. A. and Kelly, J. E., J. Polym. Sci., Polym. Chem. Edn, 1984, 22, 863.
71. Moore, J. A. and Kelly, J. E., Polymer, 1979, 20,627.
72. Moore, J. A. and Partain, E. M., Macromolecules, 1983, 16, 338.
73. Hirai, II., J. Macromol. Sci., Chem., Part A, 1984, 21, 1165.
74. Roudet.. J., Doctorate thesis, National Polytechnic Institute, Grenoble, France, 1987; Roudet, J.
and Gandini, A., submitted.
75. Storbeck, R. and Ballauff, M., Polymer, 1993, 34, 5003.
76. Khrouf. A., Boufi, S., El-Gharbi, R., Belgacem, N. M. and Gandini, A., Polym. Bull., 1996,37,
589.
77. Maccio, D., Thesis, University of Genova, Italy, 1988.
78. Costa, G., Gandini, A., Morinelli, S. and Valenti, B., Heterocycles, 1993, 36, 1225.
79. Vriesma, B. K. and Miniaci, F., US Patent No. 4 876 327, 1989.
80. Cai, R., Preston, J. and Samulski, E. T., Macromolecules, 1992, 25, 563.
81. Hopff, H. and Kreiger, A., Makromol. Chem., 1961, 47,93.
82. Heertjes, P. M. and Kok, G. J., Delji Propr. Rep., 1974, Al, 59.
83. Ogata, N. and Shimamura, K., Polym. J., 1975, 7, 72.
84. Russo, M., Kunststoffe, 1975, 65, 346.
85. Moore, J. A. and Bunting, W. W., Adv. Polym. Synth., 1985, 31,51.
86. Sarzhevskaya, V. P., Komev, K. A. andzmirnova-Zamkova, S. E., Ukr. Khim. Zh., 1964,30,499.
87. Mitiakoudis, A., Gandini, A. and Cheradame, H., Polym. Commun., 1985, 26, 246.
88. Mitiakoudis, A. and Gandini, A., Macromolecules, 1991, 24, 830.
89. Mihajlov, V. M. and Peeva, N., Makromol. Chem., 1968, 116, 107.
90. Moss, Ez. K., J. Cell. Plastics, 1982, 19, 240.
91. Cawse, J. L., Stanford, J. L. and Still, R. H., Makromol. Chem., 1984, 185,708.
92. Belgacem, M. N., Quillerou, J. and Gandini, A., Eur. Polym. J., 1993, 29, 1217.
93. Belgacem, M. N., Quillerou, J., Gandini, A., Roux, G. and Rivero, J., Eur. Polym. J., 1989, 25,
1125.
94. Boufi, S., Belgacem, M. N., Quillerou, J. and Gandini, A., Macromolecules, 1993, 26, 6706.
95. Boufi, S., Belgacem, M. N. and Gandini, A., Polymer, 1995, 36, 1689.
96. Koutchltarova, L. K., Magroupov, F. A. and Abdourachidov, T. R., Dokl. Akad. Nauk USSR,
1984, 6,42.
97. Holfinger, S. M., Conner, A. H. and Hill, C. G., J. Appl. Polym. Sci., 1993, 49, 337.
98. Chundury, D. and Szmant, H. H., Polym. Prepr., 1983, 24,50.
99. Pate], A. A. and Patel, S. R., Eur. Polym. J., 1983, 19,561.
100. Hui, Z. and Gandini, A., Eur. Polym. J., 1992, 28, 1461.
101. Mealares, C. and Gandini, A., Polym. Int., 1996, 40, 33.
102. Patel, A. A. and Majmudar, G. H., J. Appl. Polym. Sci., 1991, 43, 2091.
103. Patek A. A. and Majmudar, G. H., Eur. Polym. J., 199 1, 27, 89.
104. Kehayoglou, A. H. and Karayannidis, G. P., J. Macromol. Sci., Chem., 1982, Al& 237;Kehayo-
glou, A., Karayannidis, G. and Sideridou-Karayannidou, I., Makromol. Chem., 1982, 183,293.
105. Mtchin, F., Doctorate thesis, University of Paris VI, Paris, France, 1988; Mechin, F., Boileau, S.
and Gandini, A., to be published.
106. Hong, EI. H. and Weber, W. P., Polym. Bull., 1989, 22, 363.
107. Cooke, A. W. and Wagener, K. B., Macromolecules, 1991,24, 1404.
108. Hui, Z.. Doctorate thesis, National Polytechnic Institute, Grenoble, France, 1992; Hui, Z. and
Gandini, A., in preparation.
109. Kolb, K.. E. and Wilson, C. L., Chem. Commun., 1969, 272.
1 IO. Kolb, K.. E. and Black, W. A., Chem. Commun., 1969, 1119.
1374 A. GANDINI and M. N. BELGACEM

111. Gilbert, B. C., Marshall, P. D. R., Norman, R. 0. C., Pineda, N. and Williams, P. S., J. Chem. Sot.,
Perkin Trans. 2, 1981, 1392.
112. Benati, L., la Barba, N., Tiecco, M. and Tundo, A., J. Chem. Sot. (B), 1969, 1253.
113. Schuler, R. H., Laroff, G. P. and Fessenden, R. W., J. Phys. Chem., 1973,77,456.
114. Lunazzi, L., Placucci, G. and Grossi, L., J. Chem. Sot., Perkin Trans. 2, 1982, 875.
115. Burkey, T. J., Lunazzi, L. and Nazran, A. S., J. Org. Chem., 1983, 48, 3704.
116. Encina, M. V., Rivera, M. and Lissi, E. A., J. Polym. Sci., Polym. Chem. Edn, 1978, 16, 1709.
117. Butler, G. B., Badgett, J. T. and Sharbash, M., J. Mucroml. Sci., Chem., 1970, A4,51.
118. Kamo, B., Morita, I., Horie, S. and Furusawa, S., Polym. J., 1974, 6, 121.
119. Vazzana, I., Grandi, F., Hayashi, K., Munari, S. and Russo, S., Chim. Ind. (Milan), 1975, 57,
745.
120. Gaylord, N. G., Martan, M. and Deshpande, A. B., J. Polym. Sci., Polym. Chem. Edn, 1978, 16,
1527.
121. Ragab, Y. A. and Butler, G., J. Polym. Sci., Polym. Chem. Edm, 1981, 19, 1175.
122. Kuramoto, N., Suda, K. and Nagai, K., J. Polym. Sci., Polym. Chem. Edn, 1989, 27, 1713.
123. Lax-in, P. P., Musayev, U. N., Tashpulatov, Yu. T., Tillayev, R. S. and Usmanov, Kh. U, Polym.
Sci. USSR, 1964,6, 1098.
124. Fawcett, A. H., Mulemwa, J. N. and Tan, C. E., Polym. Commun., 1984,25, 300; Fawcett, A. H.,
Mulemwa, J. N. and Tan, C. E., Polymer, 1985,26, 1268.
125. Chau, C. W., Fawcett, A. H., Mulemwa, J. N., Poon, L. W., Surgenor, A. and Tan, C. E.,
Makromol. Chem., 1992, 193, 257.
126. Fawcett, A. H., Yau, T. F., Mulemwa, J. N. andTan, C. E., Br. Polym. J., 1987,19,211; Apperley,
D. C.. Cm-ran. D. M. and Fawcett, A. H., Polvmer, 1988.29, 909.
127. Oikawa, E., Aoki, F. and Katano, K., Polym. J., 1979, 11, 257.
128. Hall, H. K. and Righettini, R. F., Polym. Bull., 1986, 16,405.
129. Trifonov, A. and Panajatov, Iv., Compt. Rend. Acad. Bulg. Sci., 1957,10,301; Trifonov, A. and
Panajatov, Iv., Compt. Rend. Acad. Bulg. Sci., 1957, 10, 363.
130. Aso, C. and Tanaka, Y., Kobunshi Kogaku, 1964,21,373.
131. Andreeva, I. V. and Koton, M. M., Dokl. Akad. Nauk SSSR, 1956, 110, 75; Koton, M. M.,
J. Polym. Sci., 1958, 30, 331.
132. Goutiere, G. and GolC, J., Bull. Sot. Chim. Fr., 1965, 162.
133. Simek, I. and Hanus, M., Chem. Zvesti, 1960, 14, 124.
134. Aso, C., Kunitake, T., Tanaka, Y. and Miyazaki, H., Kobunshi Kogaku, 1967,24, 187; Chem.
Abstr., 1968, 68, 222752.
135. Mishina, A., Nippon Nogeikagaku Kaishi, 1960, 34, 649; Chem. Abstr., 1963, 58, 14109.
136. Kamenar, S., Simek, I. and Regensbogenova, E., Chem. Zvesti, 1960, 14,581.
137. Reikhsfel’d, V. O., Trudy Vsesoyuz. Soveshchaniya, Riga, 1955, 355; Chem. Abstr., 1959, 53,
11871~.
138. Alvarez, R., Gandini, A. and Martinez, R., J. Polym. Sci., Polym. Lett. Edn, 1975, 13, 385.
139. Dillingham, K. A., Ph.D. thesis, University of Lancaster, U.K., 1994.
140. Trumbo, D. L., Suzuki, T. and Harwood, H. J., Polym. Prepr., 1983,24,367.
141. Hernandez, C. and Gandini, A., unpublished results.
142. Bachman, G. B. and Heizey, L. V., J. Am. Chem. Sot., 1949, 71, 1985.
143. Palyakova, A. M., Korshak, V. V. and Lipatnikov, N. A., Vyssokomol. Soedin., 1962, 4, 334.
144. Bachman, G. B., Filar, L. J., Finholt, R. W. and Heizey, L. V., Znd. Eng. Chem., 1951, 43,
997.
145. Mikhailov, M., Bubevska, K. and Berlin, A. A., Compt. Rend. Acad. Bulg. Sci., 1966, 19,807.
146. Mikhailov, M. and Bubevska, K., Makromol. Chem., 1968, 117,80.
147. Berlin, A. A., Bubevska, K. and Mikhailov, M., Zzv. Akad. Nauk SSSR, Ser. Chim., 1966, 943.
148. Bubevska, K., Berlin, A. A. and Mikhailov, M., Vyssokmol. Soedin., 1967, 9B, 309.
149. Bevington, J. C. and Harris, D. O., Polym. Lett., 1967, 5,799.
150. Peniche, C., Zaldivar, D., Bulay, A. and Roman, J. S., Polym. Degrad. Stab., 1993, 40,287.
151. Zaldivar, D., Peniche, C., Bulay, A. and Roman, J. S., J. Polym. Sci., Polym. Chem. Edn, 1993,31,
625.
152. Zaldivar, D., Peniche, C., Bulay, A. and Roman, J. S., Polymer, 1992, 33,4625.
FURANS IN POLYMER CHEMISTRY 1375

153. Zaldivalr, D., Davidenko, N., Peniche, C., Gallardo, A., Sartre, R. and Roman, J. S., Polymer,
1994, 35, 2390.
154. Peniche, C., Zaldivar, D., Bulay, A. and Roman, J. S., J. Appl. Polym. Sci., 1993, 50, 2121.
155. Zaldivabr, D., Peniche, C., Gallardo, A. and Roman, J. S., Biomuteriuls, 1993, 14, 1073.
156. Peniche, C., Zaldivar, D., Gallardo, A. and Roman, J. S., J. Appl. Polym. Sci., 1994, 54,959.
157. Hopff, H. and Lussi, H., Makromol. Chem., 1956, 18/19,227.
158. Hardy, G. and Szita, J., Actu Chim. Acad. Sci. Hung., 1958, 15, 338.
159. Hardy, G., Actu Chim. Acad. Sci. Hung., 1958, 17, 121.
160. Morita, I., Kato, Y., Kamo, B. and Furusawa, S., Kuyugaku Rikug. Kiyo, 1973, 16, 139; Chem.
Abstr., 1974, 81, 106028.
161. Rieumont, J. and Espinosa, R., Revistu Ciencius Quimicus, 1983, 14,293.
162. Salon, M. C., Amri, H. and Gandini, A., Polym. Commun., 1990, 31, 210.
163. Lu, Y.-H. and Hsu, C.-S., Polym. Prepr., 1995, 36 (2), 184.
164. Borrows, E. T., Haward, R. N., Porges, J. and Street, J., J. Appl. Chem., 1955, 5, 379.
165. Hemandez, C., Galego, N., Llerandi, N. and Gandini, A., Revista Cenic, 1974, 5, 217.
166. Gandini, A. and Rieumont, J., paper presented at 1st International Symposium on Radical Poly-
merization, Santa Margherita, Ligure, 17-22 May 1987.
167. Rieumont, J. and Femandez, M. E., Revista Ciencias Quimicus, 1983, 14, 285.
168. Rieumont, J. and Zaldo, A., Revistu Ciencias Quimicas, 1983, 14, 297.
169. Rieumont, J. and Davidenko, N., Revistu Cubana Quimica, 1985, 1,71.
170. Rieumont, J., Armas, L. M., Davidenko, N. and Paz, J. A., Revistu Cubana Quimica, 1985,1, 107.
171. Rieumont, J. and Davidenko, N., Actu Polym., 1986, 37, 297.
172. Rieumont, J., Davidenko, N. and Paz, J. A., Acta Polym., 1987, 38, 21.
173. Rieumont, J., Vega, R., Davidenko, N. and Paz, J. A., Eur. Polym. J., 1988,24,909.
174. Davidenko, N., Zaldivar, D., Peniche, C., Same, R. and Roman, J. S., Polym. Chem. Ed., 1996,
34, 2759.
175. Hashihama, Y., Shono, T. and Ishigaki, A., Technol. Rep. Osaka Univ., 1963, 13,481.
176. Korshak, V. V., Sultanov, A. S. and Abduvaliyev, A. A., Uzb. Khim. Zh., 1959, 4, 39; Chem.
Abstr., 1960, 54, 12642.
177. Topchiev, A. V., Goldfarb, Y. Y. and Krentsel, V. A., Vyssokomol. Soedin., 1963, 3, 870.
178. Khaidarov, K. F., Abduvaliyev, A. A. and Sultanov, A. S., Vyssokomol. Soedin., 1963, 5, 1012.
179. Khaidarov, K. F., Abduvaliyev, A. A. and Sultanov, A. S., Uzb. Khim. Zh., 1964, 4, 65; Chem.
Abstr., 1965, 62, 6683.
180. Ishigaki, A., Shono, T. and Hachihama, Y., Kogyo Kagaku Zusshi, 1963.33, 119; Chem. Abstr.,
1963,59,4047.
181. Ishigaki, A. and Shono, T., Bull. Chem. Jpn, 1974, 47, 1467.
182. Armour, M., Davies, A. G., Upadhyay, J. and Wassermann, A., J. Polym. Sci., 1967, A15, 1527.
183. Kresta, J. and Livingston, H. K., Polym. Lett., 1970, 8,795.
184. Barr, D. A. and Rose, J. B., J. Chem. Sot., 1954,3766.
185. Livingston, H. K., Senkus, R., Hsien, T.-T. and Kresta, J., Makromol. Chem., 1967, 161, 101.
186. Shindo. K. and Ishikawa, S., J. Chem. Sot. Jpn, Chem. Ind. Chem., 1977,2,245.
187. Lamb, B. S. and Kovacic, P., J. Polym. Sci., Polym. Chem. Edn, 1980, 18, 2423.
188. Koton, M. M., J. Appl. Chem. USSR (Engl. Trunsl.), 1941, 14, 181.
189. Hachihama, Y. and Imito, M., J. Sot. Chem. Ind. Jpn (Suppl.), 1942,45, 190; Chem. Abstr., 1950,
44, 9721f.
190. Aso, C. and Tanaka, Y., Kobunshi Kahaku, 1964,24,373; Chem. Abstr., 1965,62,9293a.
191. Stoicescu, C. and Dimonie, M., Rev. Roum. Chim., 1968, 13, 109.
192. Alvarez, R., Gandini, A. and Martinez, R., J. Polym. Sci., Polym. Lett. Edn, 1975, 13, 385.
193. Gandini, A. and Martinez, R., J. Polym. Sci., Polym. Symp., 1976, 13,79.
194. Gandini, A., Martinez, R. and Sanchez, R., Revista Cenic Ciencias Fisicas, 1979, 10, 13; Chem.
Abstr., 1980, 93, 221136.
195. Gandini, A. and Martinez, R., Makromol. Chem., 1983, 184, 1189.
196. Gandini, A. and Martinez, R., Revista Ciencias Quimicas, 1985, 16, 275.
197. Martinez, R. and Gandini, A., Revista Cenic Ciencius Quimicus, 1987, 18, 100.
198. Martinez, R. and Gandini, A., Revista Cubana Quimicas, 1988, 4, 39.
1376 A. GANDINI and M. N. BELGACEM

199. Martinez, R., Alvarez, R. and Gandini, A., Acta Polym., 1989, 40,613.
200. Plesch, P. H., J. Chem. Sot., 1953, 1662.
201. Garreau, H. and Marechal, E., Eur. Polym. J., 1973,9,263.
202. Rooney, J. M., Makromol. Chem., 1978,179,2419.
203. Martinez, R., Alvarez, R. and de La Guardia, C., Polym. Bull., 1995,34,57.
204. Reijendam, J. W. and Janssen, M. J., Tetrahedron, 1970,26, 1303; Reijendam, J. W., Heeres, G. J.
and Janssen, M. J., Tetrahedron, 1970, 26, 1291.
205. Gandini, A. and Plesch, P. H., Eur. Polym. J., 1968, 4, 55.
206. Prosser, H. J. and Young, R. N., Eur. Polym. J., 1972, 8, 879.
207. Salon, M. C., Doctorate thesis, National Polytechnic Institute, Grenoble, France, 1984; Salon, M.
C. and Gandini, A., to be published.
208. Rieumont, J., Doctorate thesis, University of Havana, Cuba, 1975.
209. Ivin, K. S. and Busfield, W. K., in Encyclopedia of Polymer Science and Engineering,
Vol. 12, eds. H. F. Mark, N. M. Bikales, C. G. Overberger and G. Menges. Wiley, New York,
1988, p. 555.
210. Raff, R., Cook, J. L. and Ettling, B. V., J. Polym. Sci., 1965, 3A, 35 11.
211. Gandini, A. and Rieumont, J., Br. Polym. J., 1977, 28.
212. Korshak, V. V., Cejtlin, G. M., Chomutov, V. A. and Gonzales, Ch.P., Acta Polym., 1980, 31,
587.
213. Korshak, V. V., Tseitlin, G. M., Khomutov, V. A., Gonsales, Kh.P. and Atrushkevich, A. A.,
Polym. Sci. USSR, 1979, 21, 6 1.
214. Korshak, V. V., Solov’eva, L. K., Gribova, I. A., Makina, L. B., Krasnov, A. P., Petrovskii, P. V.,
Komarova, L. I., Kulakov, V. V. and Kabanova, E. L., Polym. Sci. USSR, 1980, 22, 2734.
215. Makina, L. B., Korshak, V. V., Solovyeva, L. K., Gribova, I. A., Krasnov, A. P., Petrovskii, P. V.,
Komarova, L. I. and Kabanova, Ye. L., Polym. Sci. USSR, 1981,23,2229.
216. Feringa, B. L., Hulst, R., Rikers, R. and Brandsma, L., Synthesis, 1988, 316; Comins, D. L. and
Killpack, M. O., J. Org. Chem., 1987, 52, 104; Ramanathan, V. and Levine, R., J. Am. Chem.
Sot., 1962, 27, 12316.
217. Gout&e, G., Leonetti, J.-B. and GoI& J., Compt. Rend., 1963,257,2485; Gout&e, G., Doctorate
thesis, University of Lyon, France, 1963.
218. Normant, H. and Angelo, B., Bull. Sot. Chim. (Fr.), 1961, 10, 1988.
219. Kasai, P. H. and McLeod, D. Jr., J. Am. Chem. Sot., 1973, 95,480l.
220. Andreeva, I. V. and Koton, M. M., Dokl. Akad. Nauk SSSR, 1956, 110, 75.
221. Kuwata, K., Kawazura, H. and Hirota, K., Nippon Kagoku Zasshi, 1960,81, 1770; Chem. Abstr.,
1962,56,4928.
222. Iwakura, Y. and Toda, F., J. Macromol. Sci., Chem., 1975, A9, 1063.
223. Gandini, A. and Hernandez, C. D., Polym. BULL, 1978, 1,221.
224. Natta, G., Mazzanti, G., Pregaglia, G. F. and Pozzi, G., J. Macromol. Sci., 1962, 58,
1201.
225. Aharoni, S. M. and Walsh, E. K., Macromolecules, 1979, 12, 27 1; Parodi, F., in Comprehensive
Polymer Science, Vol. 5. Pergamon Press, Oxford, 1989, p. 387.
226. Hui, Z. and Gandini, A., Polym. Bull., 1991, 26, 383.
227. Ionescu, I. V., Kenyeres-Ursu, V. and Isacescu, D. A., Buletinul Institutului Politechnic Din Iasi,
1970, 16,419.
228. Patel, A. A. and Patel, S. R., Eur. Polym. J., 1983, 19, 231.
229. Kahmamedov, T. K., Abdiev, 0. B. and Gacanov, F. D., Doklady Acad. Nauk Azerb. SSR, 1979,
35, 36.
230. Wirpsza, Z. and Wolszczak, M., Polymeri, 1977,22,7; Vinoslavskii, V. A., Kamenskii, I. V. and
Yeryshev, B. Ya., Polym. Sci. USSR, 1979, 17, 1108.
23 1. Amri, H., Belgacem, M. N. and Gandini, A., Polymer, 1996, 37,48 15.
232. Amri, H., Belgacem, M. N., Signoret, C. and Gandini, A., Polym. Int., 1996,41,427.
233. Gandini, A. and Hernandez, C. D., Makromol. Chem., Rapid Commun., 1981,2,351.
234. Razzouk, H., Bouridah, K., Gandini, A. and Cheradame, H., in Cationic Polymerization and
Related Processes, ed. E. J. Goethals. Academic Press, New York, 1984, p. 355.
235. Salon, M. C. and Gandini, A., unpublished results.
FURANS IN POLYMER CHEMISTRY 1377

236. Russel, K. E. and Vail, L. G. M. C., Can. J. Chew., 1979, 57, 2355; Hunter, B. K., et al.,
J. Polym. Sci., Polym. Chem. Edn, 1983, 21, 435; Rooney, J. M., J. Appl. Polym. Sci., 1980,
25, 1365.
237. Gandini, A., Salon, M. C., Choura, M., El-Gharbi, R., Amri, H. and Hui, Z., Mukromol. Chem.,
Macromol. Symp., 1992, 60, 165.
238. Gandini, A. and Salon, M. C., Deutsch Patent No. 8900137, 1989 (assigned to Stamicarbon).
239. Jiang, Z., Sanganeria, S. and Sen, A., J. Polym. Sci., Polym. Chem. Edn, 1994,32,841.
240. Bredas, J. L. and Chance, R. R., Conjugated Polymeric Materials: Opportunities in Electronics,
Optoelectronics and Molecular Electronics. Kluwer Academic Publishers, New York, 1990;
Skotheim, T. A., Handbook of Conducting Polymers. Marcel Dekker, New York, 1986; Feast,
W. J., Tsibouklis, J., Pouwer, K. L., Groendaal, L. and Meijr, E. W., Polymer, 1996, 37, 5017.
241. Tourillon, G. and Garnier, F., J. Electroanal. Chem., 1982, 135, 173.
242. Ohsawa, T., Kaneto, K. and Yoshino, K., Jpn J. Appl. Phys., Part 2, 1984,23,663; Yoshino, K.,
Hayashi, S. and Sugimoto, R., Jpn J. Appl. Phys., Part 2, 1984,23, 899.
243. Kang, E. T. and Neoh, K. G., Eur. Polym. J., 1987,23,7 19.
244. Tediar, F., Eur. Polym. J., 1985, 21, 3 17.
245. Downard, A. J. and Pletcher, D., J. Electroanal. Chem., 1986, 206, 147.
246. Beck, F. and Pruss, A., J. Electroanal. Chem., 1987,216, 157.
247. Zotti, G., Schiavon, G., Comisso, N., Berlin, A. and Pagani, G., Synth. Met., 1990, 36, 337.
248. Kauffmann, T. and Lexy, H., Chem. Ber., 1981, 114,3667.
249. Distefano, G., Jones, D., Gerra, M., Favaretto, L., Modelli, A. and Mengoli, G., J. Phys. Chem.,
199 1,95,9746.
250. Joshi, M. V., Helmer, C., Cava, M. P., Cain, J. L., Bakker, M. G., McKinley, A. J. and Metzger, R.
M., J. Chem. Sot. Perkin Trans. 2, 1993, 1081; Joshi, M. V., Cava, M. P., Bakker, M. G.,
McKinley, A. J., Cain, J. L. and Metzger, R. M., Synth. Met., 1993, 55-57, 948.
251. Glenis, S., Benz, M., Legoff, E., Shindler, J. L., Kannewurf, C. R. and Kanatzidis, M. G., J. Am.
Chem. Sot., 1993,115, 12519.
252. Hemandez, V., Ramirez, F., Zotti, G. and Lopez Navarette, T. J., Chem. Phys. Lett., 1992, 191,
419; Hernandez, V., Ramirez, F., Zotti, G. and Lopez Navarette, T. J., Synth. Met., 1993,55-57,
4467; Hemandez, V., Ramirez, F., Zotti, G. and Lopez Navarette, T. J., J. Chem. Phys., 1993,2,
769.
253. Carrillo, J., Sanchez de la Blanca, E., Gonzalez-Tejera M. J. and Hernandez-Fuentes, I., Chem.
Phys. Zett., 1994,229,633; Gonzalez-Tejera, M. J., Carrillo, J. and Hernandez-Fuentes, I., Anal.
Quim., 1993,89,52 1; Gonzalez-Tejera, M. J., Carrillo, J. and Hernandez-Fuentes, I., Synth. Met.,
1995,73, 153
254. Bakhshi, A. K., J. Chem. Phys., 1992, 96, 2339.
255. Otto, P., Int. J. Quantum Chem., 1994, 52, 353.
256. Jurim,ae, T., Strandberg, M., Karelson, M. and Calais, J.-L., Int. J. Quantum Chem., 1995,54,369.
257. Nishioka, Y. and Yoshino, K., Jpn J. Appl. Phys., 1990, 29, 1675.
258. Yoshino, K., Manda, Y., Takahashi, H., Nishioka, Y., Kawai, T., Ohmori, Y. and Onoda, M., J.
Appl. Phys., 1990, 68, 5976.
259. Wang, S., Yoshino, K., Tanaka, K. and Yamabe, T., J. Phys. Sot. Jpn, 1991,60,2002.
260. Reynolds, J. R., Child, A. D., Ruiz, J. P., Hong, S. Y. and Marynick, D. S., Macromolecules, 1993,
26, 2095.
261. Hong. S. Y. and Marynick, D. S., Macromolecules, 1992, 25, 3591.
262. Gandini, A. and Mealares, C., Trends Polym. Sci., 1994, 2, 127.
263. Kossmehl, G. and Yaridjanian, A., Makroml. Chem., 1981, 182, 3419.
264. Boro, I. and Hinrichsen, G., Polym. Bull., 1994, 33,471.
265. Yamada, S., Tokito, S., Tsitsui, T. and Saito, S., J. Chem. Sot., Chem. Commun., 1987, 1448.
266. Murase, I., Ohnishi, T., Noguchi, T. and Hirooka, M., Polym. Commun., 1987, 28, 229.
267. Jen, K. Y., Maxfield, M., Shacklette, L. W. and Elsenbaumer, R. L., J. Chem. Sot., Chem.
Commun., 1987, 309; Jen, K. Y., Jow, T. R. and Elsenbaumer, R. L., J. Chem. Sot., Chem.
Commun., 1987, 1113.
268. Mealalres, C., Hui, Z. and Gandini, A., Polymer, 1996, 37, 2273.
269. Mealalres, C. and Gandini, A., Eur. Polym. J., 1996,32, 1269.
1378 A. GANDINI and M. N. BELGACEM

270. De Schryver, D., Doctorate thesis, University of Antwerpen, Belgium, 1989.


271. Kossmehl, G. and Greczmiel, T., Mukromol. Chem., 1992, 193, 335; Kossmehl, G. and Grecz-
miel, T., Macromol. Chem. Phys., 1994, 195, 3655.
272. Hayakawa, Y., Nishida, M., Okumura, A., Matsul, M. and Muramatsu, H., Polym. Bull., 1982,28,
293.
273. Allcock, H. R., Dodge, J. A., Van Dyke, L. S. and Martin, C. R., Report TR-68, 1992, Order No.
AD-A249747, 56 pp. Available from: NTIS, Gov. Rep. Announce. Index (U.S.) 1992, 92(16),
Abstr. No. 243 408.
274. Mealares, C., Costa, J., Gandini, A. and Valenti, B., Acta Polym., 1997, 48, 21.
275. Braslavsky, S. and Heiklen, J., Chem. Rev., 1977, 77,473.
276. Price, D., Ratajczak, E., Sztuba, B. and Sarzynski, D., J. Photochem., 1987,37, 273.
277. Gandini, A., Parsons, J. M. and Back, R. A., Can. J. Chem., 1976, 54, 3095.
278. Feringa, B. L., Rec. Trav. Chim. Pays-Bus, 1987, 106,469.
279. Andrews, D. J. and Feast, W. J., J. Polym. Sci. Polym., Chem. Edn, 1976, 14,331.
280. Hasegawa, M., Adv. Polym. Sci., 1982, 42, 1; Dilling, W. L., Chem. Rev., 1983, 83, 1.
281. Reizer, A., Photoreactive Polymers, The Science and Technoology of Resists. John Wiley and
Sons, New York, 1989.
282. Williams, J. L. R. and Borden, D. G., Makromol. Chem., 1964, 73, 203.
283. Lahav, M. and Schmidt, G. M. J., J. Chem. Sot. (B), 1967, 239.
284. Karminski-Zamola, G. and Jakopcic, K., Croatia Chem. Acta, 1974, 46,71.
285. Tsuda, M., J. Polym. Sci., Part A-l, 1969, 7, 259.
286. Hasegawa, M., Suziki, Y., Suziki, F. and Kikuchi, S., J. Polym. Sci., Part A-l, 1969, 7, 743.
287. Rave, A., Pasternack, G., Brown, K. H. and Raldove, S. B., J. Polym. Sci., Polym. Chem. Edn,
1974,11, 1733.
288. Muller, S., Le Barny, P., Chastaing, E. and Robin, P., Molec. Eng., 1992, 2, 251.
289. Bare& V., Gandini, A. and Rousset, E., J. Photochem. Photobiol., 1997, 103, 171; Baret, V.,
Doctorate thesis, National Polytechnic Institute, Grenoble, France, 1995.
290. Carruthers, W., Cycloaddition Reaction in Organic Synthesis. Pergamon Press, Oxford, 1990.
291. Fringuelli, F. and Taticchi, A., Dienes in the Diels-Alder Reaction. John Wiley & Sons, New
York, 1990.
292. Berlin, A. A., Liogon’kii, B. I., Zapadinskii, B. I., Kazantzeva, E. A. and Stankevich, A. O.,
J. Macromol. Sci., Chem., 1977, All, 1.
293. Tesoro, G. C. and Sastri, V. R., Ind. Eng. Chem. Prod. Res. Dev., 1986, 25, 444.
294. He, X., Sastri, V. R. and Tesoro, G. C., Makromol. Chem., Rapid Commun., 1989, 9, 191.
295. Patek H. S. and Lad, B. D., Makromol. Chem., 1989,190,2055; Patel, H. S., Lad, B. D. and Vyas,
H. S., High Perf Polym., 1990, 2, 113.
296. Patek H. S. and Patel, H. D., Phosphorus, St&r, Silicon, 1993, 79, 69.
297. Blatter, K. and Schltiter, A.-D., Macromolecules, (1989,22,3506; Vogel, T., Blatter, K. and Schl
titer, A.-D., Makromol. Chem., Rapid Commun., 1989, 10, 427.
298. O’Dell, R., Ph.D. thesis, Lancaster University, U.K., 1990.
299. Lai’ta, H., Doctorate thesis, National Polytechnic Institute, Grenoble, France, 1993; GoussC, C.,
Doctorate Thesis, National Polytechnic Institute, Grenoble, France, 1997.
300. Edelmann, D. and Ritter, H., Mukromol. Chem., 1993, 194, 1183.
301. Ritter, H., Sperber, R. and Weisshuhn, C. M., Mukromol. Chem., 1993, 194, 1721.
302. Edelmann, D. and Ritter, H., Makromol. Chem., 1993, 194, 2375.
303. Reinecke, M. and Ritter, H., Makromol. Chem., 1993, 194,2385.
304. Stevens, M. and Jenkins, A., J. Polym. Sci., 1979, 17, 3675.
305. Ktlsefoglu, S. H., J. Polym. Sci., 1984, 22,255.
306. Chujo, Y., Sada, K. and Saegusa, T., Macromolecules, 1990, 23, 2636.
307. Andre, D. and Gandini, A., unpublished results.
308. Laita, H., Boufi, S. and Gandini, A., European Polym. J. (in press).
309. Gousse, C. and Gandini, A., Eur. Polym. J. 1997, 33, 667.
310. GoussC, C., Hodge, P. and Gandini, A., paper presented at International Conference on Recent
Advances in Polymer Synthesis, University of York, U.K., July 1996.
311. Ivin, K. J., Olejin Methatesis. Academic Press, London, 1983.
FURANS IN POLYMER CHEMISTRY 1379

312. Novae, B. M. and Grubs, R. H., J. Am. Chem. Sot., 1988,110,960; Novae, B. M. and Grubs, R. H.,
J. Am. Chem. Sot., 1988,110, 7542.
313. Hamilton, J. G., Marquess, D. G., O’Neill, T. J. and Rooney, J. J., J. Chem. Sot., Chem. Commun.,
1990, 119.
314. Feast, W. J. and Harisson, D. B., Polym. Bull., 1991, 25, 343; Feast, W. J. and Harisson, D. B.,
Polymer, 1991, 32, 558.
315. Hillm:yer, M. A., Lepetit, C., McGrath, D. V., Novae, B. M. and Grubbs, R. H., Macromolecules,
1992,25,3345.
3 16. Seehof, N. and Risse, W., Makromol. Chem., Rapid Commun., 1991, 12, 107.
3 17. Bazan, G. C., Oskam, I. H., Cho, N.-H., Park, L. Y. and Schorck, R. R., J. Am. Chem. Sot., 1991,
113, 6,899.
318. Lu, S.-Y., Quayle, P., Heatley, F., Booth, C., Yeates, S. G. and Padget, J. C., Macromolecules,
1992,25,2692.
3 19. Viswa.nathan, T. and Jethmalani, J., J. Appl. Polym. Sci., 1993, 48, 1289.
320. Viswanathan, T. and Jethmalani, J., J. Appl. Polym. Sci., 1993,47, 1477.
321. Viswanathan, T., Gomez, F. and Jethmalani, J., J. Polym. Sci., Polym. Chem. Edn, 1994,32,2469.
322. Safir, A. L. and Novae, B. M., Macromolecules, 1993,26,4072.
323. Novae, B. M. and Safir, A. L., Polym. Prepr., 1996,37,335.
324. Safir, A. L. and Novae, B. M., Polym. Prepr., 1995, 36,227

You might also like