Vous êtes sur la page 1sur 23

Assisted Microwave Synthesis of High Molecular

Weight Poly(ArylEtherKetone)s

RAPHAL BRUNEL
CATHERINE MARESTIN1
VINCENT MARTIN
REGIS MERCIER
FREDERIC SCHIETS
Laboratoire des Matriaux Organiques Proprits Spcif iques, UMR 5041 CNRS /
Universit de Savoie, BP 24 69390 Vernaison, France
(Received 15 January 20071 accepted 2 April 2007)

Abstract: A microwave-assisted aromatic nucleophilic substitution reaction has been used to synthesize
high molecular weight Poly(ArylEtherKetone)s (PAEKs) within very short reaction times and through a
highly simplified process. The influence of different parameters, (namely the microwave power, solvent
nature, or reaction time) on the polymer molecular weights were studied in order to optimize the poly-
merization conditions. The polymers thus obtained were characterized and compared with their analogues
synthesized by conventional thermal polycondensation. This work underlines the potential of microwave-
assisted aromatic polymer synthesis.

Key Words: Microwave, poly(ArylEtherKetone) synthesis, polycondensation

1. INTRODUCTION

Since its discovery in 1945, microwave irradiation technology has been used for heating
and drying materials. However, its first use in the field of organic chemistry synthesis was
not until 1979 [1] and it has been further investigated since 1986 [2, 3]. Since this period,
tremendous interest has been devoted to assisted microwave reactions as illustrated by the
exponential rise in the number of publications related to this field [4] and to the wide diver-
sity of organic reactions that are successfully performed [510] (Suzuki coupling, Claisen
rearrangements, nucleophilic substitutions, Fries rearrangements, oxidations. . . ). Indeed,
many advantages are associated with assisted microwave irradiation, such as very rapid
reaction, high yield, great selectivity and even environmentally-friendly experimental con-
ditions. Controversy has however appeared concerning specific or non thermal effects,
called microwave effects [11]. Since then, many publications [1214] have compared

High Performance Polymers, 20: 185207, 2008 DOI:10.1177/0954008307079617


1
12008 SAGE Publications

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


186 R. BRUNEL ET AL.

different reaction types in analogous experimental conditions, by both conventional ther-


mal and assisted microwave methods. Up to now, no general and systematic evidence of
specific effects has yet been reported. In many cases, (in homogeneous systems in partic-
ular), the most probable reason for microwave efficiency relies on the quasi-instantaneous
temperature profile and the direct heating of the reaction medium which are known as
thermal effects [15].
Whatever the origin of the effectiveness, the microwave-assisted microwave synthesis
seems particularly interesting for the preparation of polymers. Recent reviews [16, 17]
report on the growing interest of such processes for a wide range of polymerization path-
ways, such as ring opening polymerization [18, 19], living radical polymerization [20],
CC coupling polymerization [10, 21, 22], step growth polymerization [23, 24]. High
performance phenylethynyl imide oligomers were even recently successfully prepared by
microwave-assisted solid-state synthesis and produced at a 300 g scale by NASA [25].
As reported by Loupy et al. [8] assisted microwave ether synthesis via a nucleophilic
substitution reaction is particularly efficient. Whereas some work has been reported on the
experimental conditions of arylether organic compounds (in the presence of a phase trans-
fer catalysis, in solventless conditions [26, 27]. . . ), relatively few data are available de-
scribing the synthesis of polyaryl(ether)s [27, 28]. To the best of our knowledge, no work
related to the microwave-assisted synthesis of fully aromatic Poly(ArylEtherKetone)s
(PAEKs) has yet been reported.
PAEKs are known to offer excellent performances such as good solvent resistance,
high thermo-oxidative stability and good mechanical properties. For these reasons, they
are used in many different fields such as the electronics and aerospace industries. More
recently, PAEKs modified by the introduction of sulfonic groups along the backbone were
investigated as proton conductive membranes for fuel cells [29, 30].
This work demonstrates that it is possible to obtain high molecular weight PAEKs
using microwave irradiation and underlines the potential of such a polymerization proce-
dure.

2. EXPERIMENTAL

2.1. Materials

2,2-Bis(4-hydroxyphenyl)propane (Bisphenol A) (Sigma-Aldrich) was recrystallized


from toluene, 4,42 -hexafluoroisopropylidene diphenol (Bisphenol AF) and 4,42 -dichloro-
diphenylsulfone were purified by sublimation, hydroquinone and 4,42 -dichlorobenzo-
phenone were, respectively, purified by crystallisation in water and acetonitrile. All other
reagents (4-hydroxy,42 -methoxybenzene, 4-fluorophenol, (4-chlorophenyl)phenylsulfone,
4-fluorobenzophenone, phenol), solvents and bases chlorobenzene, N-methyl pyrro-
lidinone (NMP) (SDS), dimethyl formamide (DMF), dimethyl sulfoxide (DMSO), potas-
sium carbonate (K2 CO3 ) and sodium carbonate (Na2 CO3 )- were used as received. 4,42 -
difluorobenzophenone (ACROS) was purified by crystallization from a petroleum ether/

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 187

ethyl acetate mixture. 4,42 -Bis(4-fluorobenzoyl)diphenyl ether was prepared according to


the procedure described by Hergenrother et al. [31].

2.2. Characterization

2.2.1. Nuclear magnetic resonance


1
H and 13 C nuclear magnetic resonance (NMR) spectra were recorded on a Brcker
250 MHz spectrometer using CDCl3 as solvent, at room temperature, with tetramethyl
silane (TMS) as reference. In 13 C assignments, 3 , 33 and 333 stands for exchangeable car-
bons. 19 F spectra were recorded on a multi-probe Brcker 400 MHz, using CFCl3 as
reference.

2.2.2. Size exclusion chromatography

The polymer molecular weights were determined by size exclusion chromatography (SEC)
in DMF + 0.05 mol L41 LiBr, at 705 C, and calibrated with polystyrene standards.

2.2.3. Viscosity measurements

Inherent viscosities were measured at 305 C from a 0.5 dl g41 solution of polymer in NMP,
with an Ubbelohde capillary viscosimeter.

2.2.4. Differential scanning calorimetry

The glass transition temperature (T g ) of the polymers as determined by differential scan-


ning calorimetry (DSC) with a Mettler-Toledo DSC822e apparatus at a heating rate of
55 C min41 .

2.3. Synthesis of PAEK by conventional thermal polycondensation

Typically, in a 50-mL, three-necked, round-bottom flask equipped with a nitrogen in-


let and a mechanical stirrer, 3 g (8.92 6 1043 mol) of 4,42 -hexafluoroisopropylidene
diphenol, 2.958 g (21.4 6 1043 mol) of K2 CO3 , and 1.94 g (8.92 6 1043 mol) of 4-
42 -difluorobenzophenone were refluxed 2 h in a mixture of 19 mL of NMP and 10 mL
of toluene. The water released during the bisphenate formation was stripped off by
azeotropic distillation using a Dean Stark trap. The temperature was then increased up
to 1605 C during different reaction times in order to complete the polymerization. After
cooling, the mixture was poured into methanol. The polymer was chopped into pieces and
then collected by filtration, washed with methanol and dried under vacuum at 805 C for
4 h.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


188 R. BRUNEL ET AL.

2.4. Synthesis of PAEK by assisted microwave polycondensation

Polymerizations under microwave radiation were achieved in a multi-mode Milestone mi-


crowave oven.
In a single-neck, 50-mL, round-bottom flask equipped with a mechanical stirrer,
1.94 g (8.92 6 1043 mol) of 4-42 -difluorobenzophenone, 3 g (8.92 6 1043 mol) of 4,42 -
hexafluoroisopropylidene diphenol and 2.958 g (21.4 6 1043 mol) of potassium carbonate
were added to 19 mL of NMP. The reaction medium was then submitted to microwave ir-
radiation with various intensities and for different reaction times. The viscous solution
thus obtained was cooled down before precipitation in methanol. The white fibrous poly-
mer was chopped into small pieces and collected by filtration, washed with methanol and
dried under vacuum, at 805 C for 4 h.

2.5. Membrane preparation

Thin films were prepared by the casting solution method. A 1520 wt.% polymer solution
in NMP previously filtered on a 45 1m filter was cast onto a glass substrate. The films
thus obtained were heated at 505 C overnight and then successively treated 1 h at 805 C, 1 h
at 1305 C and finally 2 h at 1805 C.

2.6. Model compound synthesis (scheme 1)

2.6.1. Synthesis of para (4-methoxyphenoxy)phenylsulfone (1)

In a single-neck, round-bottom flask equipped with a magnetic stirrer and a condenser


were placed 20 g (6.96 6 1042 mol) of dichlorodiphenylsulfone, 17.72 g (0.14 mol) of
paramethoxyphenol1 23.1 g (0.167 mol) of potassium carbonate and 150 mL of DMAC.
The reaction mixture was submitted to microwave irradiation (300 W for 3 h). When the
reaction mixture was completed (as witnessed by thin layer chromatography), the medium
was poured into water. After a while, a white powder could be filtered and further washed
with methanol. Para (4-methoxyphenoxy)phenylsulfone (1) was dried under vacuum at
505 C (T f = 935 C) and was obtained with 91.3% yield.
1
H NMR (CDCl3 ): 2 = 3.81 (s, H9 )1 6.976.93 (m, H3 , H6 and H7 ), 7.82 (d,H2 ,
J 233 = 8.8 Hz). 13 C NMR (CDCl3 ): 2 = 55.6 (C9 ), 115.1 (C33 ), 116.8 (C63 )1 121.8 (C7 )1
129.6 (C2 )1 135 (C1 )1 148 (C5 )1 157 (C8 )1 162.8 (C4 ).

2.6.2. Synthesis of para(4-hydroxyphenoxy)phenylsulfone (2)

In a 125-mL teflon reactor were added 4 g (8.6 6 1043 mol) of para(4-methoxyphenoxy)


phenylsulfone and 12 g (10.2 6 1042 mol) of pyridinium hydrochloride. The reactor was
submitted to microwave irradiation for 10 min at 300 W. The temperature was monitored
by an optical fiber probe (up to 1905 C) dipped in the reaction medium and the pressure
was recorded during the reaction (up to 1.3 bars). After cooling the medium to room

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 189

Scheme 1. Structures of model compounds.

temperature, the medium was precipitated into water and washed intensively with water.
Para(4-hydroxyphenoxy)phenylsulfone (2) was obtained as a pure white powder with 85%
yield.
1
H NMR (DMSOd6 ): 2 = 6.8 (d, H7 ), 6.937.02 (m, H3 and H6 ), 7.85 (d, H2 ).
13
C NMR (DMSOd6 ): 2 = 116.7 (C33 ), 116.8 (C63 )1 122 (C7 )1 129.8 (C2 )1 134.5 (C1 )1
146.1 (C5 )1 155 (C8 )1 162.7 (C4 ).

2.6.3. Synthesis of 1,4-bis-(4-benzenecarboxyphenoxy)benzene (3)

In a 50 mL, single-neck, round-bottom flask equipped with a condenser were placed


2 g (9.98 6 1043 mol) of 4-fluorobenzophenone, 0.5365 g (4.87 6 1043 mol) of hy-
droquinone and 1.615 g (1.16 6 1042 mol) of K2 CO3 in 10 mL of NMP. The reaction
mixture was heated at 1705 C overnight. After precipitation in water and several washings
with methanol, a light grey powder was obtained with 91.2% yield and characterized as
pure 1,4-bis-(4-benzenecarboxyphenoxy)benzene (3) by NMR.
1
H NMR (CDCl3 /ATF(1/1)): 2 = 7.15 (d, H4 ), 7.24 (s, H1 ), 7.56 (t, H9 3 J 938 = 7.45 Hz),
7.71 (t,H10 3 J 9310 = 7.45 Hz), 7.8 (d,H8 3 J 839 = 8.35 Hz), 7.92 (d,H5 3 J 435 = 8.9 Hz).
13
C NMR (CDCl3 /ATF(1/1)): 2 = 117.6 (C43 ), 122.8(C13 ), 129.1(C533 ), 131.0(C833 ), 134.4
(C10 ,C933 ), 137.0 (C7 ), 152.5 (C2 ),164.1 (C3 ), 202.2 (C=O).

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


190 R. BRUNEL ET AL.

2.6.4. Synthesis of 1,4-bis(4-benzenesulfonylphenoxy)benzene (4)

The same procedure as above was used with 6 g (2.37 6 1042 mol) of 4-chlorophenyl)
phenylsulfone, 1.275 g (1.16 6 1042 mol) of hydroquinone, 3.839 g (2.77 6 1042 mol)
of K2 CO3 and 28 mL of DMSO. The resulting beige powder was identified as 1,4-bis(4-
benzenesulfonylphenoxy)benzene (4) by NMR.
1
H NMR (CDCl3 /ATF(1/1)): 2 = 7.12 (d, H4 ), 7.14 (s, H1 ), 7.557.71 (m, H9 and
H10 ) 7.927.97 (m, H 5 and H8 ). 13 C NMR (CDCl3 /ATF(1/1)): 2 = 118.9 (C4 )1 123.50
(C1 )1 128.3 (C9 )1 130.8 (C 83 ),131.1 (C53 )1 133.8 (C6 )1 135.4 (C7 )1 140.8 (C10 ), 152.9 (C2 ),
164.3 (C3 ).

2.6.5. Synthesis of 4,42 bis[4-(4-benzenecarbonylphenoxy)phenoxy]benzenesulfone (5)

A similar procedure as for the synthesis of 1,4-bis-(4-benzenecarboxyphenoxy)benzene


(3) was used, with 0.164 g (8.12 6 1044 mol) of 4-fluorobenzophenone, 0.174 g (4 6
1044 mol) of (2), 0.133 g (9.4 6 1044 mol) of K2 CO3 in 1.3 mL of NMP. After isolation,
4,42 -bis[4-(4-benzenecarbonylphenoxy)phenoxy]benzenesulfone (5) was characterized by
NMR.
1
H NMR (CDCl3 /ATF (1/1)): 2 = 7.127.22 (m, H3 , H6 , H7 , H10 ), 7.56 (t, H16 ), 7.71
(dd, H17 ), 7.81 (dd, H15 ), 7.907.97 (m, H11 and H12 ). 13 C NMR (CDCl3 /ATF (1/1)):
2 = 118.1 (C33 ), 118.9 (C103 ), 123.4 (C6 and C7 ), 129.7 (C133 ), 130.8 (C17 ) 131.7 (C113 ),
131.7 (C12 ), 134.1 (C1 ), 135 (C2 ), 137.5 (C14 ), 152.6 (C533 ), 153.3 (C833 ), 164.3 (C4333 ),
164.6 (C9333 ).

2.6.6. Synthesis of (1-phenoxy)-4-(benzenesulfonyl)benzene (6)

2.6.6.1. Conventional thermal procedure.


In a 50-mL, round-bottom flask equipped with a condenser were placed 5 g (1498 6
1042 mol) of 4-(chlorophenyl)phenylsulfone, 1.8620 g (149861042 mol) of phenol, 3.28 g
(243861042 mol) of potassium carbonate and 16 mL of DMSO. The reaction mixture was
heated at 1605 C for 4 h. 5.4 g of (1-phenoxy)-4-(benzenesulfonyl)benzene (6) was iso-
lated by precipitation in water and filtration (yield 88%).

2.6.6.2. Procedure under microwave.


In a 50-mL, round-bottom flask were placed 5 g (1.98 6 1042 mol) of 4-(chlorophenyl)
phenylsulfone, 1.8620 g (1.98 6 1042 mol) of phenol, 3.28 g (2.38 6 1042 mol) of potas-
sium carbonate and 16 mL of DMSO. The reaction mixture was submitted to 300 W for
2 h. The reaction temperature (measured by InfraRed) reached 1485 C. (1-phenoxy)-4-
(benzenesulfonyl)benzene (6) was isolated by precipitation in water and filtration (yield
88%).
1
H NMR (DMSO-d6 ): 2 = 7.087.16 (m, H3 , H6 ), 7.247.3 (tt, H1 ), 7.47 (t, H2 ),
7.67.72 (m, H11 and H12 ), 7.938.01 (m, H7 and H10 ).

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 191

2.6.7. Synthesis of (1-parafluorophenoxy)-4-(benzenesulfonyl)benzene (7)

In a 50-mL, round-bottom flask, 2 g (7.91 6 1043 mol) of 4-chlorophenylsulfone, 0.887 g


(7.91 6 1043 mol) of 4-fluorophenol, 1.31 g (9.48 6 1043 mol) of potassium carbonate
were suspended in 15 mL of DMSO. After 90 min irradiation with microwaves (300 W)
(1-parafluorophenoxy)-4-(benzenesulfonyl)benzene (7) was isolated by precipitation in
water and dried at 805 C.
1
H NMR (DMSO-d6 ): 2 = 7.10 (d, H3 , J 233 = 5 Hz), 7.187.33 (m, H2 , H6 ), 7.597.70
(m, H11 , H12 ), 8.0 (d, H7 , H10 , J 637 = 17.5 Hz)1 19 F NMR: 117.2 ppm.

2.6.8. Synthesis of (4-phenoxy)benzophenone (8)

In a round-bottom flask equipped with a condenser were introduced 5 g (2.49 6 1042 mol)
of 4-fluorobenzophenone, 2.35 g (2.49 6 1042 mol) of phenol, 4.14 g (2.99 6 1042 mol)
of potassium carbonate and 20 mL of DMSO. The reaction mixture was heated at 1605 C
for 8 h. (4-phenoxy)benzophenone (8) was isolated by precipitation in water and filtration
(yield 85%).
1
H NMR (DMSO-d6 ): 2 = 7.1 (d, H6 , J 637 = 10 Hz), 7.12 (d, H3 , J 233 = 7.5 Hz)1 7.3 (t,
H1 , J 132 = 7.5 Hz), 7.47.6 (m, H2 and H11 ), 7.67.8 (m, H7 , H10 , H12 ).

2.6.9. Synthesis of (4-fluorophenoxy)benzophenone (9)

In a 50 mL, round-bottom flask, 2 g (9.98 6 1043 mol) of 4-fluorobenzophenone, 1.119 g


(9.98 6 1043 mol) of 4-fluorophenol, 1.65 g (11.9 6 1043 mol) of potassium carbonate
were suspended in 15 mL of DMSO. After 90 min irradiation with microwaves (300 W),
4-fluorophenoxy)benzophenone (9) was isolated by precipitation in water, washed with
water and dried under vacuum. 1 H NMR (DMSO-d6 ): d = 7.067.11 (d, H3 ), 7.167.31
(m, H2 and H6 ), 7.517.57 (m, H11 ), 7.66 (tt, H12 ), 7.707.81 (m, H7 and H10 ).19 F NMR:
117.8 ppm.

2.7. Transetherif ication reaction

2.7.1. Procedure under microwave irradiation

In a 50-mL, round-bottom flask, 4 g (1.28 6 1042 mol) of (1-phenoxy)-4-(benzenesul-


fonyl)benzene (6), 1.445 g (1.28 6 1042 mol) of 4-fluorophenol, 2.13 g (1.54 6 1042 mol)
of potassium carbonate, 1.075 g (4.29 6 1043 mol) of 4-trifluorobenzophenone (as inter-
nal standard) were suspended in 16 mL of DMSO. The reaction mixture was submitted to
microwave irradiation (300 W) and samples were regularly taken from the medium and
analyzed by fluorine NMR.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


192 R. BRUNEL ET AL.

2.7.2. Under conventional heating

When the reaction was performed under conventional heating, the above-mentioned re-
actants were placed in a three-neck flask equipped with a Dean Stark trap. After 2 h at
1505 C, the azeotropic distillation was completed and the temperature was raised up to
1605 C. Samples were then regularly taken from the reaction medium and analyzed by
19
F NMR.
The above-mentioned procedures were used as well with the sulfone model (6).

2.8. Synthesis of a random polymer by conventional heating

Typically, in a 50 mL, three-necked, round-bottom flask equipped with a nitrogen inlet and
a mechanical stirrer, 1.009 g (9 6 1043 mol) of hydroquinone, 1.316 g (446 6 1043 mol)
of dichlorodiphenylsulfone, 3 g (242 6 1043 mol) of K2 CO3 , and 1 g (446 6 1043 mol)
of 4-42 -difluorobenzophenone were refluxed for 2 h in a mixture of 13 mL DMSO and
8 mL toluene. The water released during the bisphenate formation was stripped off by
azeotropic distillation using a Dean Stark trap. The temperature was then increased up to
1605 C for an additional 6 h in order to complete the polymerization. After cooling, the
mixture was poured into methanol. The polymer chopped into pieces was then collected
by filtration, washed with methanol and dried under vacuum at 805 C for 4 h.

2.9. Synthesis of an alternated copolymer by conventional heating

In a 50-mL, three-necked, round-bottom flask equipped with a nitrogen inlet and a


mechanical stirrer, 1.004 g (446 6 1043 mol) of difluorobenzophenone, 1.526 g (141 6
1042 mol) of K2 CO3 , and 2 g (446 6 1043 mol) of para(4-hydroxyphenoxy)phenylsulfone
were refluxed for 2 h in a solvent mixture of 12 mL DMSO and 8 mL toluene. The wa-
ter released during the bisphenate formation was stripped off by azeotropic distillation
using a Dean Stark trap. The temperature was then increased up to 1605 C for an addi-
tional 6 h in order to complete the polymerization. After cooling, the mixture was poured
into methanol. The polymer was filtered, chopped into pieces, washed with methanol and
dried under vacuum.

2.10. Synthesis of random and alternated PEEK by assisted microwave polycondensation

The same proportions of monomers as reported above were introduced (excepted toluene)
in a single-neck, round-bottom flask. The reaction medium was then submitted to mi-
crowave irradiation (300 W) for 40 min. The polymers thus obtained were cooled down
before precipitation in methanol. The polymers were collected by filtration, chopped into
pieces, washed with methanol and dried under vacuum, at 805 C for 4 h.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 193

Scheme 2. Synthesis of poly (aryl ether ketone)s.

3. RESULTS AND DISCUSSION

In this study, it was not the intention to compare strictly microwave-assisted polyconden-
sations with thermal polymerizations performed in the same reaction conditions because
the reaction times required for conventional and microwave-assisted polymerizations are
not compatible. Our purpose was to optimize the experimental conditions for microwave-
assisted synthesis in order to obtain high molecular weight polymers and to compare these
conditions with those necessary to synthesize high molecular weight PAEKs by thermal
conventional methods. The last section concerns the characterization of the PAEKs ob-
tained by both procedures.
Poly(aryl ether ketone)s are classically synthesized by polycondensation reactions be-
tween a biphenol and a diaryl halide [31]. In a typical procedure, a bisphenate reagent is
first of all formed in-situ, in the presence of a base (K2 CO3 , Na2 CO3 ...). An azeotropic
distillation carried out in the presence of toluene is then necessary to remove the water
produced. Many hours are usually required to obtain the bisphenate and to remove wa-
ter from the reaction medium. The temperature is then increased in order to allow the
polymerization to proceed. In order to obtain high molecular weight polymers, relatively
long reaction times are required (at least 6 to 8 h). The molecular weights of the poly-
mers thus obtained can be evaluated from viscosity measurements and SEC analysis. In
the present work, the molecular weights, measured after reacting for 6 and 12 h, were
M n56h6 = 97 200 g mol41 and M n512h6 = 111 366 g mol41 , respectively.
In order to compare the conventional thermal way and a microwave-assisted pro-
cedure, this study used well known Poly(ArylEtherKetone)s structures (scheme 2). In
microwave-assisted experiments, the monomers (biphenol and dihalide) and the base were
added to a solvent (chlorobenzene, NMP or DMSO) and submitted to microwave irradi-
ation. The reactions were performed in a one-pot synthesis, without removing the water
formed. As will be described thereafter, PAEKs having properties similar to the poly-
mers isolated through the conventional polycondensation method could be obtained in

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


194 R. BRUNEL ET AL.

Table 1. Infra red and optical fiber plateau temperatures (300 W microwave irradiation).
Solvent IR temperature Optical fiber temperature
(5 C) (5 C)
Chlorobenzene 116 119
NMP 172 200
DMSO 166 193
Chlorobenzene + K2 CO3 100 102
NMP + K2 CO3 170 200
DMSO + K2 CO3 163 186
Polymerization in chlorobenzene 90
Polymerization in NMP 160
Polymerization in DMSO 144

very short times (40 min). The influence of different experimental parameters on the final
polymer molecular weights were investigated (microwave power, nature of the solvent,
reaction time, presence of a phase transfer catalyst, solid content, monomer nature).

3.1. Inf luence of the microwave irradiation power

The final polymer molecular weight strongly depends on the microwave irradiation power.
Indeed, when a low power (100250 W) was used, only oligomers could be isolated. At
high irradiation powers (7 300 W), a strong superheating effect of the solvent was rapidly
detected and the microwave irradiation was automatically shut down for safety reasons.
From these preliminary observations, and in order to maintain a constant microwave ir-
radiation during the whole polymerization experiments, a 300 W microwave power was
chosen. In a typical experiment, and for a given solvent, temperature increases rapidly up
to a plateau value and this temperature was kept constant for the rest of the polymerization.

3.2. Inf luence of the reaction solvent

The macroscopic properties of a solvent strongly depends on its complex permittivity 8


(8 = 82 j822 ), as this parameter characterizes the ability of a dielectric medium to absorb
and store dielectric energy. More precisely, whereas the real permittivity 82 measures the
capacity of the medium to store electric charges, the dielectric loss factor 822 characterizes
the ability of the medium to dissipate energy. As a consequence, chlorobenzene, NMP and
DMSO, which are high boiling point solvents (T eb = 132, 202 and 1895 C, respectively),
can, according to their real permittivity, be ranked as low (82Chlorobenzene = 2.6), medium
(82NMP = 32.2) and high (82DMSO = 45) absorbing media [32]. In table 1 the temperatures
reached when these different solvents are submitted to microwave irradiation (300 W) are
reported. While irradiated, the solvents were stirred magnetically and the temperature was
recorded either with the external infrared probe or with an optical fiber directly placed into
the medium.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 195

As expected, when the non-polar chlorobenzene was submitted to a microwave irra-


diation, the temperature increased very slowly and stabilized at a moderate value (1165 C).
In this case, good agreement between the temperatures measured with an infrared probe
and an optical fiber was obtained. However, when a more polar solvent was used (NMP
or DMSO), the temperature profile was different. Indeed, polar molecules are known to
absorb microwave energy very well, leading thereby to high temperatures in a very short
time. As expected, the temperature increase was very rapid: NMP was heated up to 2005 C
in less than 5 min and DMSO reached 1935 C in 3 min. It is worth noticing that the final
temperature was reached in shorter times when the experiments were done in DMSO. One
possible explanation could be the higher dielectric loss of DMSO (822 = 37.125) compared
with the moderate value of NMP (822 = 8.855), and so to a better efficiency to transfer
energy into the medium. In order to study the effect induced by the presence of K2 CO3 , a
second set of experiments was realized by adding 2.958 g of K2 CO3 in 19 mL of each sol-
vent (same proportions as used for a polymerization reaction). Surprisingly, in all cases,
the final temperature was lower (variation of 2 to 165 C) than the temperature of the pure
solvent in the same reaction conditions.
The three different solvents were then used as polymerization media for the poly-
condensation of 4,42 -hexafluoroisopropylidene diphenol with 4,42 -difluorobenzophenone.
All polymerizations were performed rigorously in the same conditions (same polymer-
ization device, solid content, amount of K2 CO3 ). As the optical fiber could not be used
with our polymerization vessel, the evolution of temperature was recorded by an infrared
probe. Being aware of the limits of such measurements, these values were considered as
indicative.
When the polymerization was performed in chlorobenzene, the temperature increase
of the reaction medium was very low, as well as the maximum temperature reached
(905 C). In this case, no high molecular weight polymers were obtained, as confirmed by
the lack of viscosity increase of the solution mixtures, and no material could be isolated by
precipitation at the end of the experiment. On the other hand, when the polymerizations
were performed in NMP or DMSO, the temperature of the reaction mixture increased very
rapidly (in a few seconds) to reach 160 and 1445 C, respectively. These temperatures were
kept constant during the rest of the irradiation. In both cases, these temperatures were
lower than those reached with the pure solvent (9T = 225 C for DMSO and 9T = 125 C
for NMP).
For a given reaction time (40 min 300 W), only oligomers could be isolated from
chlorobenzene. However, higher molecular weight polymers could be successfully syn-
thesized in DMSO (94 700 g mol41 ) and to a lower extent, in NMP (53 500 g mol41 ). One
possible explanation for these observations could come from the higher dielectric loss
factor of DMSO and to a better efficiency to transfer energy to the monomers. From this
preliminary study, DMSO appeared as the best candidate to perform assisted-microwave
PAEKs.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


196 R. BRUNEL ET AL.

Figure 1. Variation of molecular weights of PAEK with microwave irradiation time ( DMSO, NMP).

3.3. Inf luence of the reaction time on molecular weight

The influence of the irradiation time on the final molecular weights of the PAEKs obtained
from the polycondensation of 4,42 -difluorobenzophenone and 4,42 -hexafluoroisopropyli-
dene diphenol was further investigated. For this purpose, different polymerizations were
performed by varying the reaction time. The polymers thus obtained were analyzed by
SEC. The results are reported on figure 1. As can be seen, the molecular weights increase
rapidly and reach a peak after a reaction time of 3040 min. From those polymers, tough
membranes could be obtained by the solvent casting method. However, extended exposure
times to microwave irradiation have been found to be detrimental to the molecular weights
of the polymers. In a similar way, a molecular weight decrease was also observed in the
case of a conventional thermal polymerization, but for reaction times longer than 12 h.

3.4. Inf luence of a phase transfer catalysis

Phase transfer catalysis (such as N-Neopentyl-4-(dialkylamino)pyridinium chloride) are


well known to promote aromatic nucleophilic substitution in apolar solvents. Moreover,
these compounds have also been shown to successfully enhance aromatic nucleophilic
substitution reaction in polar aprotic solvents such as dimethyl acetamide (DMAC) [33].
In this particular case, the better results obtained in the presence of a phase transfer cataly-
sis were attributed to an improved solubility and nucleophilicity of the bisphenate species.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 197

Table 2. Solubility of the microwave assisted synthesized polymers.

S, soluble1 pS, partially soluble1 Sw, swollen.

For this purpose, different agents (tertiobutyl ammonium bromide, methyltriphenylphos-


phonium) were tested in order to determine their influence on the molecular weights
of the final polymers as well as on the kinetics of the microwave-assisted synthesis of
PAEKs. No real improvement was obtained with these phase transfer agents. Instability
of such compounds was suspected at the high temperatures reached during the polymer-
ization. Further work should be done by using guanidinium salts as they seem particularly
efficient, according to Gao et al. [27].

3.5. Inf luence of the monomer structure

Using the above-described optimum experimental conditions, namely 40 min reaction


time and 300 W microwave irradiation power, different macromolecular structures were
synthesized, depending on the monomers used (bisphenol A, bisphenol AF, 4,42 -difluoro-
benzophenone, 4,42 -bis(4-fluorobenzoyl)diphenyl ether).
The polymer structures were analyzed by NMR. All the spectra obtained were consis-
tent with the polymer backbones envisaged. One example of a 4,42 -difluorobenzophenone
and bisphenol AF based PAEK is represented in figure 2.
All the structures (table 2) obtained were soluble or partially soluble in common or-
ganic solvents (CH2 Cl2 , THF, DMF, NMP, DMSO) and were amorphous polymers, as
confirmed by DSC analysis.
As it could be expected, the incorporation of 4,42 -hexafluoroisopropylidene diphenol
improves the solubility of the polymer when compared with their analogues synthesized
with 2,2-bis(4-hydroxyphenyl)propane. In order to compare the influence of the poly-
merization mode on the polymer properties, DSC analysis were performed on polymers,
synthesized by both methods. For the sake of a reliable comparison, polymers of similar

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


198 R. BRUNEL ET AL.

Figure 2. 1 H and 13
C NMR spectra of microwave assisted PAEK. (2 C(CF3 )2 = 64 ppm1 X :
2 CF3 = 124 ppm ).

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 199

Table 3. Glass transition temperatures of PAEKs.

, not determined because of partial insolubility reasons1 MW, microwave-assisted polycondensation1 Th,
thermal conventional polycondensation.

molecular weights were chosen. The results are reported in table 3. For each polymer
structure, the glass transition temperatures determined by DSC were similar whatever the
polymerization mode used.

3.6. Synthesis and characterization of alternated and random poly(arylether ketone)s


poly(arylether sulfone)s copolymers

In order to highlight any dependence of polymer microstructure with the polymeriza-


tion activation source (either thermal or microwave), two different polyaryl(ethers) were
envisaged: a purely alternating poly(ether ether sulfone) (ether ether ketone) and its
random analogue. For this purpose, a specific biphenol (2) was synthesized by aromatic
nucleophilic substitution between dichlorodiphenylsulfone and paramethoxy phenol and
subsequent demethylation [34] (scheme 3). These reactions achieved under microwaves
led to a polymer grade monomer with 77.6 % yield.
The copolymerization of dichlorodiphenylsufone with hydroquinone (HQ) and
difluorobenzophenone could end up with the formation of the three following sequences:
SulfoneEther-Ether-Ketone (SEEK), Sulfone-Ether-Ether-Sulfone (SEES) and Ketone-
Ether-Ether-Ketone(KEEK) (scheme 4), leading to a random microstructure. In this
case, the proportion of each sequence type should depend on the relative reactivity of the

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


200 R. BRUNEL ET AL.

Scheme 3. Synthesis of biphenol (2).

Scheme 4. Random and alternating microstructure of poly(arylether)s.

different monomers involved. On the other hand, the condensation of biphenol (2) with
either dichlorobenzophenone (DCBz) or difluorobenzophenone (DFBz) was expected to
lead to purely alternated Sulfone-Ether-Ether-Ketone(SEEK) sequences.
Both copolymer types were therefore synthesized by conventional thermal procedure
and under microwave irradiation.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 201

13
Figure 3. C NMR [150154 ppm] of (a) alternating copolymer1 (b) random copolymer.

3.7. Inf luence of the polymerization mode on the microstructure

Whatever the polymerization mode, the copolymers obtained from biphenol (2) were sol-
uble in a large range of common organic solvents (DMSO, DMF, THF, CHCl3 , cyclopen-
tanone, 1,2-dichloroethane,
-butyrolactone...) whereas their random analogues were not.
This first macroscopic observation was consistent with the expected different microstruc-
tures of these two copolymers. From a structural point of view, both copolymer types
were characterized by 13 C NMR (figure 3). As foreseen, some differences appear in the

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


202 R. BRUNEL ET AL.

Scheme 5. Structure of the model compounds representative of the different sequences of the
copolymers.

Table 4. Copolymers microstructure characterization by 13 C NMR.


SEEK (and KEES) KEEK SEES
(%) (%) (%) (dL g41 )
HQ/DFBz/DCDPS MW 46 29 25 0.79
HQ/DFBz/DCDPS Th 44 30 26 0.6
DFBz/Biphenol(2) MW 100 / / 0.68
DFBz/Biphenol(2) Th 100 / / 0.57
DCBz/Biphenol(2) MW 70 14 16 0.43
DCBz/Biphenol(2) Th 78 8 14 0.17
MW, microwave-assisted polycondensation1 Th, thermal conventional polycondensation.

copolymers spectra, especially in the [152154 ppm] range which characterizes the qua-
ternary carbons of the hydroquinone unit. More precisely, only two peaks corresponding
to a SEEK or KEES sequence were observed for the alternated copolymer whereas four
peaks were observed for the random copolymers, due to the existing SEEK, SEES and
KEEK sequences.
The precise assignment of the different carbons was elucidated according to the rig-
orous characterization of three model molecules (3, 4, 5) (scheme 5).
On the basis of quantitative 13 C NMR experiments, the microstructure of different
copolymers was elucidated. The results are reported in table 4. Comparing the microstruc-
ture of two copolymers, obtained either by the conventional method or under microwave
irradiation, gave no significant differences in the relative proportions of the various se-
quence types observed, which suggests that no specific microstructure could be induced
by polymerization under microwaves.
Considering the copolymers obtained by copolymerization of biphenol (2), when the
comonomer used was difluorobenzophenone (DFBz), both polymerization methods were
successful in producing high viscosity polymers, and the microstructure was found to be

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 203

Scheme 6. Model reaction to investigate the occurrence of transetherifications by 19 F NMR.

purely alternating in both cases. However, when dichlorobenzophenone (DCBz) was used
instead of its difluoride analogue, only oligomers could be isolated through a conventional
thermal process (as witnessed by the very low viscosity value). When the same reaction
was performed under microwaves, higher molecular weight polymers could be synthe-
sized (sufficient to obtain self-sustainable films). Such a result shows to some extent the
interest of the polycondensation under microwaves. From a structural point of view, the
13
C NMR analysis of the DCBz / Biphenol (2) polymers showed a microstructure different
from the one expected. Indeed, only 70 and 78%, of purely alternating sequences were
present in the conventionally prepared oligomers and in the microwave-assisted copoly-
mer, respectively. To explain such an observation, some transetherification side reactions
were suspected. To confirm this hypothesis, one investigation was made using a model
compound (scheme 6) in experimental conditions identical to the one used during the
polymerization reactions.
As represented in scheme 6, in the case of a reaction performed under thermal heating,
parafluorophenol dissapears from the reaction medium whereas a new fluorinated com-
pound identified as (4-fluorophenoxy)benzophenone) appears. The respective proportions
of each fluorine containing species were determined by quantitative 19 F NMR analysis.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


204 R. BRUNEL ET AL.

Figure 4. Behavior of the transetherification reaction from model compounds.

In order to compare the behaviour of ether ether sulfone or ether ether ketone se-
quences towards transetherification reactions, the former experiment was either realized
on sulfone and ketone model structures and under either conventional thermal heating or
microwave irradiation. The results are reported in figure 4.
From these results, it appears that, whatever the polymerization process, a transetheri-
fication reaction occurs, with sulfone sequences being slightly more sensitive to transe-
therification than ketones. From a kinetic point of view, transetherifications are much
more rapid when the experiment is performed under microwave irradiation than in a con-
ventional thermal way.
These observations suggest that when the reactive difluorobenzophenone is involved,
the aromatic nucleophilic substitution is the main reaction, which is consistent with the
purely alternated structure observed. However, if a less reactive monomer (dichloroben-
zophenone) is used, a transetherification reaction can compete with the aromatic nucle-
ophilic substitution. As a consequence, the resulting polymer microstructure loses its
purely alternated character.
In all cases, the microstructure of the copolymers was identical whatever the polymer-
ization mode.
Despite the elucidation of the polymerization mechanism, no straighforward explana-
tion for the decrease of the polymer molecular weights for long reaction times could be
established. As reported earlier for aromatic nucleophilic substitution at high temperature
(3005 C) [35] as well as at 1605 C [36] in a conventional thermal way, the nucleophilic

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 205

Table 5. Influence of the base on the variation of molecular weight.


Reaction time (min) 20 40 60 90 120
Na2 CO3 Mn (g mol41 ) 28400 28970 32200 31300 25600
(dl/g) 0.31 0.31 0.33 0.31 0.3
K2 CO3 Mn (g mol41 ) 55970 94700 75200 57100 51200
(dl/g) 0.38 0.55 0.57 0.39 0.36

behavior of KF released during the course of the polymerization could be responsible


for ether linkage scissions. To confirm this hypothesis, 1 g of a preliminary synthesized
polyarylether ( = 0.64 dL g41 ) was dissolved in DMSO and the exact amount of KF
that would have been produced in the course of the polymerization reaction was added.
After submitting this mixture to microwave irradiation (300 W 60 min), the medium
was precipitated and the viscosity of the polymer collected was further measured. The
low value obtained ( = 0.14 dL g41 ) indicate a strong molecular weight decrease and
was thus attributed to the presence of KF. In order to avoid the formation of this species
during the polymerization reaction, and as NaF is supposed to have a lower nucleophilic
power, Na2 CO3 was used as a base. In this case, the sodium salt of the phenate is firstly
produced and NaF is formed in the course of the polymerization instead of KF. In such
cases, no important chain scissions were observed, even after long reaction times (up to
4 h) (table 5). However, molecular weights values were systematically lower than those
measured for the polymers synthesized in the presence of potassium carbonate. This last
phenomenon was attributed to the lower reactivity of the sodium phenate compared to the
potassium phenate in the polymerization reaction.

4. CONCLUSION

Microwave-assisted polycondensation is a very efficient method for producing PAEKs


in very short reaction times. By optimizing the reaction conditions, 40 min microwave-
assisted polymerizations were sufficient to produce polymers with high molecular weights,
whereas 6 to 8 h were necessary to reach comparable values by a conventional thermal
polymerization. Polar solvents seem, however, necessary to obtain polymers with high
molecular weights. The polymers obtained by both methods present similar properties
and characteristics (macroscopic as well as microscopic). The microwave-assisted reac-
tion is therefore particularly attractive as the polymerization process is greatly simplified
and opens up new perspectives for a wide range of aromatic and heterocyclic polymers.
Acknowledgement. A. Waton and F. Boisson from Rseau des Polymristes Lyonnais are gratefully acknowl-
edged for the 2 3 F NMR analysis.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


206 R. BRUNEL ET AL.

NOTE

1. Author to whom correspondence should be addressed: e-mail: marestin@lmops.cnrs.fr

REFERENCES
[1] Gourdenne, A., Maassarani, A.-H., Monchaux, P., Aussudre, S. and Thourel, L. (1979). Cross-linking
of Thermosetting Resins by Microwave Heating: Quantitative Approach. Polym. Prep. (ACS Div Polym.
Chem.), 20: 471.
[2] Gedye, R., Smith, F., Westaway, K., Ali, H., Baldisera, L. and Laberge, L., et al. (1986). The Use of
Microwave Ovens for Rapid Organic Synthesis. Tetrahedron Lett., 27(3): 279282.
[3] Giguere, R.J., Bray, T.L., Duncan, S.M. and Majetich, G. (1986). Application of Commercial Microwave
Ovens to Organic Synthesis. Tetrahedron Lett., 27(41): 49454948.
[4] Lidstrm, P., Tierney, J., Wathey, B. and Westman, J. (2001). Microwave Assisted Organic Synthesis a
Review. Tetrahedron, 57: 92259283.
[5] Namboodiri, V.V. and Varma, R.S. (2001). Microwave-accelerated Suzuki Cross-coupling Reaction in
Polyethylene Glycol (PEG). Green Chem., 3: 146148.
[6] Varma, R.S. (1999). Solvent Free Organic Syntheses using Supported Reagents and Microwave Irradia-
tion. Green Chem., 1, 4355.
[7] Moghaddam, F.M, Ghaffarzadeh, M. and Abdi-Oskoui, S.H. (1999). Tandem-fries Reaction Conjugate
Addition under Microwave Irradiation in Dry Media. One-pot synthesis of flavanones. J. Chem. Res. (S).,
574575.
[8] Loupy, A. (2002). Microwaves in Organic Synthesis. Wiley-VCH.
[9] Varma, R.S., Saini, R.K. and Meshram, H.M. (1997). Selective Oxidation of Sulfide to Sulfoxides and Sul-
fones by Microwave Thermolysis on Wet Silica-Supported Sodium Periodates. Tetrahedron Lett., 38(37):
65256528.
[10] Melucci, M., Barbarella, G. and Sotgiu, G. (2002). Solvent-free Microwave Assisted Synthesis of Thio-
phene Oligomers via Suzuki Coupling. J. Org. Chem., 67: 88878894.
[11] Berlan, J., Giboreau P., Lefeuvre, S. and Marchand, C. (1991). Synthse organique sous champ micro-
ondes: premier exemple dactivation spcifique en phase homogne. Tetrahedron Lett., 32(21): 2363
2366.
[12] Raner, K.D, Strauss, C.R., Vyskoc, F. and Mokbel, L. (1993). A Comparison of Reaction Kinetics Ob-
served under Microwave Irradiation and Conventional Heating. J. Org. Chem., 58: 950953.
[13] Perreux, L. and Loupy, A.A. (2001). Tentative Rationalization of Microwave Effects in Organic Synthesis
According to the Reaction Medium, and Mechanistic Considerations. Tetrahedron, 57: 91999223.
[14] Chaouchi, M., Loupy, A., Marque, S. and Petit, A. (2002). Solvent-Free Microwave-Assisted Aromatic
Nucleophilic Substitution-Synthesis of Aromatic Ethers. Eur. J. Org. Chem., 12781283.
[15] Strauss, C.R. and Trainor, R.W. (1995). Invited Review. Developments in Microwave-Assisted Organic
Chemistry. Aust. J. Chem., 48: 16651692.
[16] Wiesbrock, F., Hoogenboom, R. and Schubert, U.S. (2004). Microwave-Assisted Polymer Synthesis:
State-of- the-Art and Future Perspectives. Macromol. Rapid Commun., 25: 17391764.
[17] Zong, L., Zhou, S., Sgriccia, N., Hawley, M.C. and Kempel, L.C. (2003). A Review of Microwave-
Assisted Polymer Chemistry (MAPC). J. Microwave Power Electromag. Energy, 38(1): 4974.
[18] Liao, L.Q, Liu, L.J, Zhang, C., He, F. and Zhuo, R.X. (2003). Heating Characteristics and Polymerization
of 8-Caprolactone under Microwave Irradiation. J. Appl. Polym. Sci., 90(10): 26572664.
[19] Yu, Z.J. and Liu, L.J. (2004). Effect of Microwave Energy on Chain Propagation of Poly(8-caprolactone)
in Benzoic Acid-initiated Ring Opening Polymerization of 8-caprolactone. Eur. Polym. J., 40: 22132220.
[20] Zhang, H. and Schubert, U.S. (2004). Monomode Microwave-Assisted Atom Transfer Radical Polymer-
ization. Macromol. Rapid Commun., 25: 12251230.
[21] Tierney, S., Heeney, M. and McCulloch, I. (2005). Microwave Assisted Synthesis of Polythiophenes via
the Stille Coupling. Synth. Metals., 148: 195198.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015


HIGH MOLECULAR WEIGHT POLY(ARYLETHERKETONE)S 207

[22] Nehls, B.S., Asawapirom, U., Fldner, S., Preis, E., Farrell, T. and Scherf, U. (2004). Semiconducting
Polymers via Microwave-Assisted Suzuki and Stille Cross-Coupling Reactions. Adv. Funct. Mater., 14(4):
352356.
[23] Lewis, D.A., Summers, J.D., Ward, T.C. and McGrath, J.E. (1992). Accelerated Imidization Reactions
Using Microwave Radiation. J. Appl. Polym. Sci., 30: 16471653.
[24] Yeganeh, H., Tamami, B. and Ghazi, I. (2004). A Novel Direct Method for Preparation of Aromatic
Polyimides via Microwave-assisted Polycondensation of Aromatic Dianhydrides and Diisocyanates. Eur.
Polym. J., 40: 20592064.
[25] Smith, J.G., Connell, J.W., Li, C.-J., Wu, W. and Criss, J.M. (2006). Microwave Synthesis of Phenylethynyl
Imide Oligomers: Neat Resin and Composite Properties of PETI-298. High Perform. Polym., 18: 341354.
[26] Bogdal, D., Pielichowski, J. and Boron, A. (1998). New Synthetic Method of Aromatic Ethers under
Microwave Irradiation in Dry Media. Synth. Commun., 28(16): 30293039.
[27] Gao, C., Zhang, S., Gao, L. and Ding, M. (2004). Microwave-Assisted Synthesis of High-Molecular-
Weight Poly(ether imide)s by Phase-Transfer Catalysis. J. Appl. Polym. Sci., 92: 24152419.
[28] Hurduc, N, Abdelylah, D., Buisine, J.-M., Decock, P. and Surpateanu, G. (1997). Microwave Effects in
the Synthesis of Polyethers by Phase Transfer Catalysis. Eur. Polym. J., 33(2): 187190.
[29] Xing, P., Robertson, G.P., Guiver, M.D., Mikhlailenko, S.D. and Kaliaguine, S. (2005). Synthesis and
Characterization of Poly(aryl ether ketone) Copolymers containing (hexafluoroisopropylidene)-diphenol
Moiety as Proton Exchange Membrane Materials. Polymer, 46: 32573263.
[30] Wang, F., Chen, T. and Xu, J. (1998). Sodium Sulfonate-functionalized poly(ether ether ketone)s. Macro-
mol. Chem. Phys., 199(7): 14211426.
[31] Hergenrother, P.M., Jensen, B.J., Havens, S.J. (1988). Poly(arylene ethers). Polymer, 29: 358369.
[32] Hayes, B.L. Microwave Synthesis-Chemistry at the Speed of Light. Matthews, North Carolina.
[33] Hoffmann U., Klapper M, Mllen K. (2002). Phase-transfer catalysed synthesis of poly(ether ketone)s.
(1993). Polym Bull., 30(5): 4818.
[34] Kulkarni, P.P., Kadam, A.J., Mane, R.B., Desai, U.V. and Wadgaonkar, P.P. (1999). Demethylation of
Methyl Aryl Ethers using Pyridine Hydrochloride in Solvent Free Conditions under Microwave Irradiation.
J. Chem. Res. (S)., 394395.
[35] Fukawa, I., Tanabe, T. and Dozono, T. (1991). Preparation of Aromatic Poly(ether ketones) from an
Aromatic Dihalide and Sodium Carbonate. Macromolecules, 24: 38383844.
[36] Hoffmann, U., Helmer-Metzmann, F., Klapper, M. and Mllen, K. (1994). Poly(ether ketones) by Fluo-
rides Catalyst Systems. Macromolecules, 27: 35753579.

Downloaded from hip.sagepub.com at UNIV OF PITTSBURGH on March 11, 2015

Vous aimerez peut-être aussi