You are on page 1of 7

Article

pubs.acs.org/Macromolecules

Mechanistic Pathways for the Polymerization of Methylol-Functional


Benzoxazine Monomers
Mohamed Baqar,†,§ Tarek Agag,‡,⊥ Rongzhi Huang,‡ Joaõ Maia,‡ Syed Qutubuddin,*,†,‡
and Hatsuo Ishida‡

Department of Chemical Engineering, Case Western Reserve University, Cleveland, Ohio 44106, United States

Department of Macromolecular Science and Engineering, Case Western Reserve University, Cleveland, Ohio 44106, United States

ABSTRACT: The polymerization mechanism of methylol-


functional benzoxazine monomers is reported using a series of
monofunctional benzoxazine monomers synthesized via a
condensation reaction of ortho-, meta-, or para-methylol−
phenol, aniline, and paraformaldehyde following the traditional
route of benzoxazine synthesis. A phenol/aniline-type
monofunctional benzoxazine monomer has been synthesized
as a control. The structures of the synthesized monomers have
been confirmed by 1H NMR and FT-IR. The polymerization
behavior of methylol monomers is studied by DSC and shows
an exothermic peak associated with condensation reaction of methylol groups and ring-opening polymerization of benzoxazine at
a lower temperature range than the control monomer. The presence of methylol group accelerates the ring-opening
polymerization to give the ascending order of para-, meta-, and ortho-positions in comparison to the unfunctionalized monomer.
Furthermore, rheological measurements show that the position of methylol group relative to benzoxazine structure plays a
significant role in accelerating the polymerization.

■ INTRODUCTION
Benzoxazine resins are a new class of thermosetting polymers
more phenolic groups are formed, which promote an
autocatalytic polymerization process.6a,b
Wang and Ishida7 showed that monomers synthesized using
that are characterized by several unique properties such as near-
unsubstituted phenol exhibited variations of the ring-opening
zero shrinkage or volumetric expansion upon polymerization,
polymerization temperature. A more detailed study by Andreu
considerable molecule-design flexibility, low water absorption
et al.,8 who synthesized several substituted 3-phenyl-3,4-
and flammability, low surface energy, and high glass transition dihydro-2H-1,3-benzoxazine monomers using substituent
temperature.1 Additionally, the polymerization of benzoxazine phenol and aromatic amine, showed that increasing the
monomers takes place through thermally accelerated ring- electron-withdrawing in the para-position of phenol exhibited
opening mechanism without any added initiator or catalyst.2 a noticeable decrease in the polymerization temperature. This is
These numerous unusual properties of benzoxazine resins are attributed to the acidic phenol species that are formed due to
associated with the existence of inter- and intramolecular the electron-withdrawing groups. On the other hand, increasing
hydrogen bonds in the network structure.3a−d Therefore, in the electron-withdrawing in the para-position of the phenyl
order to understand the structure−property relationships of the substituent monomers showed an increase of the polymer-
polymer, it is of great importance to know the mechanistic ization temperature due to the destabilization of the
pathways of polymerization. propagating iminium intermediates as proposed by McDonagh
Although the polymerization chemistry of benzoxazine and Smith.9a,b However, for the electron-donating substituents
monomers still remains rather poorly understood, the high no notable effect on the polymerization was observed regardless
basicity of benzoxazine monomers, which is attributed to both of the position of the substituent.
the oxygen and the nitrogen of the oxazine ring by Lewis The effect of functionalization and copolymerization of
definition, suggests that the ring-opening polymerization of carboxylic group into benzoxazine structure has been
benzoxazine monomers proceeds through a cationic mecha- studied10a−c and a significant decrease in polymerization
nism.4a−c During the polymerization, the monomer comes to temperature was observed. Polybenzoxazine chains preferen-
equilibrium with the corresponding zwitterionic intermediate tially form intramolecular 6-membered hydrogen bond,11a−i
which eventually forms the polymer as depicted in Scheme 1. which tends to slow the propagation reaction.12a,b This
Several fundamental studies for polymerization mechanisms
were conducted to understand the polymerization pathways of Received: September 18, 2012
benzoxazine monomers.5a−c These investigations suggest that, Revised: September 21, 2012
as the polymerization of benzoxazine monomers progresses, Published: October 1, 2012

© 2012 American Chemical Society 8119 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125
Macromolecules


Article

Scheme 1. Mechanism of Ring-Opening Polymerization of RESULTS AND DISCUSSION


Benzoxazine Monomers Preparation of Monofunctional Benzoxazine Mono-
mers. A successful synthesis of methylol monofunctional
benzoxazine monomers has been achieved using aniline,
paraformaldehyde, and hydroxybenzyl alcohols (ortho-, meta-
and para-isomers) as well as phenol (P) for comparison as
indicated in Scheme 2. The monomers are designated as

Scheme 2. Preparation of Monofunctional Benzoxazine


Monomers

hydrogen bonding by COOH group to the newly formed


phenolic OH groups competes with the intramolecular 6-
membered ring formation between the phenolic and amine
groups,13 hence the retarding effect on chain propagation by (oHBA-a, mHBA-a, and pHBA-a) in which the smaller italic
the intramolecular hydrogen bonding. letters represent the position of methylol group into
Kiskan et al.14a,b synthesized hydroxyethyl terminated ether benzoxazine structure, either in ortho, meta or para,
chain-functional benzoxazine monomers and reported the respectively. P-a represents a phenol/aniline type monofunc-
reduced polymerization temperature of these monomers in tional benzoxazine monomer synthesized as a control with no
comparison to ordinary unfunctional benzoxazine monomers. methylol group in the structure.
Kudoh et al.15 subsequently studied the mechanistic aspects of The structures of the monomers were confirmed using 1H
the polymerization and the role of hydroxylethyl group in NMR and FT-IR spectra. In the 1H NMR spectra, as depicted
activating the ring-opening of hydroxyethyl functional benzox- in Figure 1, the typical resonances attributed to the benzoxazine
azine monomers. They found that the hydroxyethyl monomer structure, Ar−CH2−N− and −O−CH2−N−, for pHBA-a,
polymerized much faster than a similar structure of other N- mHBA-a, oHBA-a and P-a are observed at 4.62, 5.36; 4.64, 5.37;
alkylbenzoxazines, which is attributed to the intramolecular 4.66, 5.41; and 4.74, 5.48 ppm, respectively. Also, the 1H NMR
reaction of hydroxyl group with cationic moieties of the spectra confirm the presence of methylol group (−CH2OH)
zwitterionic intermediates formed by the ring-opening reaction from the resonance of −CH2−O of methylol group at 4.55,
4.56, and 4.64 ppm for pHBA-a, mHBA-a, and oHBA-a,
of benzoxazine to afford a 5-membered cyclic N,O-acetal.
respectively.
Sudo et al.16 showed that increasing the bulkiness of the
There are a number of infrared absorption peaks, highlighted
substituent N-alkyl group on to benzoxazine monomers leads
in Figure 2, that are used to verify the formation of oxazine
to a decrease in the polymerization rate due to the release of a rings in each monomer. For example, the characteristic out-of-
larger amount of N-alkylimine as a volatile compound. Oie et plane absorption modes of benzene with an attached oxazine
al.17 reported that N-allylbenzoxazine monomers exhibit a faster ring are located at 947, 950, 945, and 946 cm−1 for the pHBA-a,
rate of polymerization than N-(n-propyl)benzoxazines due to mHBA-a, oHBA-a, and P-a, respectively.19 Furthermore, the
the presence of allyl group on nitrogen atom. Consequently, presence of the benzoxazine ring aromatic ether in the
the intramolecular interaction between the cationic moieties monomers is indicated by an absorbance peak centered in
and electron-rich carbon−carbon double bond of this group the range of 1035−1032 cm−1 due to the C−O−C symmetric
prompted the formation of the zwitterionic intermediates. stretching mode. Also, the peak between 1242 and 1226 cm−1
In continuation of our observed reduction of the ring- for the asymmetric stretching modes confirms the presence of
opening polymerization temperature of methylol benzoxazine the oxazine ring in the monomer structure.20a,b Peaks
monomers compared to the unfunctionalized monomers,18a−c characteristic of asymmetric trisubstituted benzene that appear
this paper is aimed at studying the role of incorporating in the methylol monomers between 1505 and 1500 cm−1
methylol groups into benzoxazine structure. Furthermore, the confirm the incorporation of methylol group into benzoxazine
polymerization mechanism of benzoxazine monomers with monomers. Furthermore, the presence of methylol group in the
various isomeric methylol positions is proposed. synthesized monomers is confirmed through the very broad
8120 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125
Macromolecules Article

Figure 1. 1H NMR spectra of benzoxazine monomers.

Figure 2. FT-IR spectra of benzoxazine monomers.

absorption peak between 3340 and 3320 cm−1 due to the OH


stretching mode.21a,b
Effect of Methylol Group on the Polymerization Figure 3. DSC thermograms of benzoxazine monomers.
Behavior of Benzoxazine Monomers. The polymerization
profiles of methylol benzoxazine monomers were studied by Table 1. Results of the DSC Analysis of Benzoxazine
DSC as depicted in Figure 3 and the results are summarized in Monomers
Table 1. The thermograms show that the onset of the ring- heat of polymerization
opening polymerization of unfunctionalized benzoxazine monomer onset temp (°C) max temp (°C) (J/g)
monomer started at 237 °C with its maximum centered at P-a 237 255 336
255 °C. However, the onset of polymerization is shifted in the oHBA-a 151 196 201
presence of methylol group to as low as 151, 183, and 205 °C mHBA-a 183 214 203
for oHBA-a, mHBA-a, and pHBA-a, respectively. In addition, pHBA-a 205 231 241
the peaks are significantly broadened and the maxima are
shifted to temperatures of 196, 214, and 231 °C for oHBA-a, From the DSC results, the values of the heat of polymer-
mHBA-a, and pHBA-a, respectively. Moreover, the methylol- ization for the functionalized monomers were lower than the
functional benzoxazine monomers show lower values for the control, showing clearly that another thermal event is involved
heat of polymerization compared to the control. For example, in the ring-opening reaction of methylol benzoxazines. Since
the heat of polymerization values were 201, 203, and 241 J/g the presence of methylol groups is the main feature of these
for oHBA-a, mHBA-a, and pHBA-a, respectively. However, P-a, monomers, one possible thermal event may be attributed to the
which has no methylol groups, exhibits the higher exotherm of release of water as a byproduct from the polymerization of
336 J/g. methylol groups similar to traditional phenolics.22a,b The
8121 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125
Macromolecules Article

occurrence of methylol condensation reaction together with closed molecular structure in comparison to open structure that
ring-opening polymerization of methylol benzoxazine mono- tends to extend.
mers has been confirmed elsewhere.18a Since both the storage and loss moduli eventually attain
Chemorheological Studies of Monomers. Viscoelastic maximum values and remain at the plateaus, the following
properties during the polymerization of each monomer were rheological fractional conversion is defined assuming reaction is
evaluated during isothermal polymerization at 140 °C. A complete:23
qualitative comparison of the rates of polymerization for the
monomers is determined by the increase of the viscoelastic G′(t )
α=
moduli as shown in Figure 4. The figure shows that methylol- G′(α) (1)
where G′(t) is the storage modulus at time t and G′(α) is the
maximum modulus at the end of polymerization. The resultant
time conversion plots are shown in Figure 5. The fractional
conversion increases rapidly immediately after the reaction is
prompted, due to the double catalytic effects of methylol and
phenol as stated above.

Figure 4. Storage (G′) and loss (G″) modulus as a function of time at


140 °C for oHBA-a (□, ○),mHBA-a (◊, ◁), pHBA-a (△, ▽), and P-
a (▷, ○) monomers.

functional benzoxazine monomers react faster than the control


(P-a) because of the lack of initiating species other than the
impurity level phenolic structures. The concentration of the
Figure 5. Isothermal fractional conversion as a function of time for
phenolic structures builds gradually and autocatalytic process
methylol monomers.
starts, accelerating the rate of polymerization. In the polymer-
ization of P-a, G′ did not increase for 45 min, and then
increased suddenly. However, in the case of methylol Figure 6 shows a number of highlighted infrared bands,
monomers, the onset of polymerization occurred at much which can be used to verify the ring-opening reaction and the
shorter times: oHBA-a exhibits the quickest onset of subsequent polymerization of oxazine rings in each monomer.
approximately 6 min, followed by mHBA-a at 20 min and For example, the monomers, pHBA-a, mHBA-a, oHBA-a, and
pHBA-a at 32 min. This is because of the catalytic effect of the P-a show a decrease in the characteristic absorption bands of
methylol groups, in addition to the aforementioned autocata- the benzene ring to which oxazine is attached that are located at
lytic effect of the phenolic structure produced by the ring- 947, 950, 945, and 946 cm−1, respectively. In addition, the
opening reaction of the oxazine rings. absorption bands located in the range of 1242−1226 cm−1 due
The DSC technique is often used to study the polymer- to the symmetric stretching of C−O−C are decreased
ization behavior of benzoxazine monomers through ring- confirming the ring-opening polymerization. Furthermore,
opening polymerization. The observed values of the curing methylol monomers show a decrease in the peaks characteristic
peak of the DSC thermogram indicate the local reactivity of the of asymmetric trisubstituted benzene between 1505 and 1500
monomers toward ring-opening. The chemorheological results cm−1 while the appearance of new peaks around 1490 cm−1 due
are interpreted with emphasis on the onset of the increase in to the formation of tetra-substituted benzene confirms the
storage modulus which also represents the reactivity of the polymerization of methylol benzoxazine monomers.
monomer. Therefore, the DSC peaks and G′ results are Both DSC and rheological analysis show that the methylol
consistent in their trend: (ortho, meta, para). However, the G′ group accelerates the autocatalytic ring-opening polymerization
and G″ crossover represents not the reactivity but the of benzoxazine. The remarkable difference in reactivity of
formation of the global infinite network. Hence the network methylol monomers over the unfunctionalized benzoxazine
formation does not have to follow the trend observed in monomer suggests a mechanism where the methylol group has
exothermic peak. In other words, the local reactions detected by an important role to promote the ring-opening polymerization.
DSC do not necessarily have to agree with the global network In Scheme 3, only the o-methylol mechanisms for the cross-
formation. This is the case if the local structure tends to form linking reaction are shown as an example. In this proposal, the
8122 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125
Macromolecules Article

moiety.4b,9a,14−16 Another type of zwitterionic intermediate (B)


is a tautomer of the intermediate (A) which is considered as a
carbocationic. Since the methylol monomers have a higher
possibility for hydrogen bonding, and the intermediate (A + B)
would be unstable because of the high concentration of OH
ions. Therefore, by the ionization of phenol and as nucleophilic
phenolic nuclei are strengthened, the methylene bridge
formation is favored by means of the formation of phenoxide
ions. The zwitterionic intermediate (A and B) then proceed
through different pathways where water is a byproduct of the
methylol condensation toward the formation of methylene
linkages. In addition, the resulting structure has an iminium and
phenoxide moieties, of which reactions can contribute to
increase number of cross-linking points and eventually a
network structure of polybenzoxazine. The proposed mecha-
nims in the current study is fundamentally different from the
mechanisms proposed by Kudoh et al. for hydroxyethyl
functional benzoxazine since methylol group is incapable of
forming the intermediate structure proposed by those
authors.15
Moreover, the DSC thermograms show that the ring-opening
polymerization is shifted to the lower curing temperature as the
methylol group is closer to the oxygen atom in the cyclic
Figure 6. FT-IR spectra of unpolymerized and 140 °C polymerized benzoxazine. The results support the proposed mechanism
monomers. where, in the case of o-methylol, the resonance of the
benzoxazine ring is affected by the methylol to form the
monomer reaches equilibrium with the corresponding zwitter- hydrogen bonding. This ring activates the oxazine ring to open
ionic intermediate (A) which has a phenoxide and an iminium at lower temperature. However, the mechanisms propose that

Scheme 3. Proposed Polymerization Mechanism of Methylol Benzoxazine Monomers

8123 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125


Macromolecules Article

benzoxazine ring is less affected by methylol in the case of meta- stretching of Ar−O−C), 1033 (symmetric stretching of C−O−C),
and para-monomers as supported by the DSC thermograms. 950 (out-of-plane vibration, benzene ring to which oxazine is


attached).
1
H NMR spectra, δH (300 MHz, CDCl3, TMS, ppm): 4.56 (s,
CONCLUSIONS −CH2−O), 4.64 (s, C−CH2−N), 5.37 (s, N−CH2−O−), 6.79−7.30
A series of methylol functional benzoxazine monomers with (m, Ar).
different hydroxybenzyl alcohol isomers were successfully Preparation of (3-Phenyl-3,4-dihydro-2H-benzo[e][1,3]-
synthesized. The DSC results show that exothermic peaks oxazin-6-yl)methanol [abbreviated as pHBA-a]. In a 500 mL
bottom-rounded flask, pHBA (14.95 g, 120 mmol), aniline (11.18 g,
due to condensation reaction of methylol groups and ring- 120 mmol), and paraformaldehyde (7.52 g, 250 mmol) were mixed
opening polymerization of benzoxazine are 231, 214, and 196 together and refluxed for 6 h in toluene (135 mL). The product was
°C for monomers that methylol group placed on para-, meta-, concentrated using a vacuum evaporator and redissolved in chloroform
and ortho-position, respectively. However, the exothermic peak followed by base-washed then once with water. The product was then
of unfunctionalized monomer shows a higher value of 255 °C. dried over anhydrous sodium sulfate and recrystallized from
The rheological study indicates that the onset of polymerization chloroform to yield a white product (yield: 21.15 g, 73%).
occurs at much shorter times of 6, 20, and 32 min for oHBA-a, IR spectra (KBr, cm−1): 3340 (stretching of OH of methylol), 1500
mHBA-a, and pHBA-a compared to the unfunctionalized (stretching of trisubstituted benzene ring), 1228 (asymmetric
monomer that takes 45 min. The highest reactivity of the stretching of Ar−O−C), 1032 (asymmetric stretching of C−O−C),
947 (out-of-plane vibration, benzene ring to which oxazine is
methylol monomers is attributed to the catalytic effect of the attached).
methylol group on the ring-opening due to intramolecular 1
H NMR spectra, δH (300 MHz, CDCl3, TMS, ppm): 4.55 (s,
hydrogen bonding between the methylol and the oxygen in the −CH2−OH), 4.62 (s, C−CH2−N), 5.36 (s, N−CH2−O−), 6.80−
benzoxazine ring as proposed in the polymerization mecha- 7.28 (m, Ar).
nism. Characterization and Measurements. Proton nuclear magnetic


resonance (1H NMR) spectra were acquired on a Varian Oxford
EXPERIMENTAL SECTION AS300 at a proton frequency of 300 MHz using an average number of
transients of 64. A relaxation time of 10 s was used for the integrated
Materials. 2-Hydroxybenzyl alcohol (oHBA) (99%), 3-hydrox- intensity determination of 1H NMR spectra. Deuterated chloroform
ybenzyl alcohol (mHBA) (99%), 4-hydroxybenzyl alcohol (pHBA) (CDCl3) was used to obtain the spectra with tetramethylsilane (TMS)
(98%), phenol (98%), and aniline (99%) were obtained from Sigma− as an internal standard. Fourier transform infrared (FT-IR) spectra
Aldrich and used as-received. Paraformaldehyde (96%) was obtained were acquired on a Bomem Michelson MB100 which was equipped
from Acros Organics, USA. Toluene, ethyl acetate, chloroform, with a deuterated triglycine sulfate (DTGS) detector at a resolution of
hexanes (a mixture of isomers), and 1,4 dioxane were obtained from 4 cm−1 with 32 coadditions. The spectra were taken by casting a thin
Fisher and used as received. film onto a KBr plate. Thermal analysis of the samples was performed
Preparation of 3-Phenyl-3,4-dihydro-2H-benzo[e][1,3]- via differential scanning calorimetry (DSC) using TA Instruments
oxazine [abbreviated as P-a]. P-a was prepared from phenol, DSC model 2920 with temperature ramped at 10 °C/min and a
aniline and paraformaldehyde following the reported method.8 IR nitrogen flow rate of 65 mL/min. All samples were crimped in
spectra (KBr, cm−1): 1230 (asymmetric stretching of Ar−O−C), 1035 hermetic aluminum pans with lids. The evolution and comparison of
(symmetric stretching of C−O−C), 946 (out-of-plane vibration, rheological properties during polymerization for each of the different
benzene ring to which oxazine is attached). benzoxazine monomers were performed utilizing an Anton Paar
1
H NMR spectra, δH (300 MHz, CDCl3, TMS, ppm): 4.74 (s, C− Rheometer (Model Physica MCR 501) with disposable parallel upper
CH2−N), 5.48 (s, N−CH2−O−). 6.99−7.44 (m, Ar). and lower plates, measuring 25 mm and 50 mm in diameter,
Preparation of (3-Phenyl-3,4-dihydro-2H-benzo[e][1,3]- respectively. Small amplitude oscillatory shear time sweep experiments
oxazin-8-yl)methanol [abbreviated as oHBA-a]. Into a 100 mL
over temperatures at 140 °C were performed using a constant
bottom-rounded flask were mixed oHBA (3.73 g, 30 mmol), aniline
frequency of 10 rad/s for all experiments, but continuously increasing
(2.8 g, 30 mmol), and paraformaldehyde (1.9 g, 60 mmol) and the
stress between a minimum of 1 Pa and a maximum of 400 Pa.
mixture refluxed in 1,4-dioxane (35 mL) for 4 days. The product was
Although the actual stress ramp varied per experiment, it was always
filtered and then concentrated using a rotary evaporator. The product
set to a range of values that kept the fluids’ response in the linear
was then redissolved in ethyl acetate and base-washed followed by
viscoelastic regime, while providing a high enough strain to provide
once with water. The organic layer was then dried over sodium sulfate
reproducible results.


anhydrous, followed by vacuum evaporation to afford viscous oily
product (yield: 3.71 g, 51%).
IR spectra (KBr, cm−1): 3320 (stretching of OH of methylol), 1505 AUTHOR INFORMATION
(stretching of trisubstituted benzene ring), 1226 (asymmetric Corresponding Author
stretching of Ar−O−C), 1032 (symmetric stretching of C−O−C), *E-mail: sxq@case.edu.
945 (out-of-plane vibration, benzene ring to which oxazine is
attached). Notes
1
H NMR spectra, δH (300 MHz, CDCl3, TMS, ppm): 4.64 (s, The authors declare no competing financial interest.
−CH2−OH), 4.66 (s, C−CH2−N), 5.41 (s, N−CH2−O−), 6.90− §
On leave from Azzaytuna University, Libya.
7.29 (m, Ar). ⊥
On leave from Tanta University, Tanta, Egypt.


Preparation of (3-Phenyl-3,4-dihydro-2H-benzo[e][1,3]-
oxazin-7-yl)methanol [abbreviated as mHBA-a]. In a 100 mL
bottom-rounded flask, mHBA (7.52 g, 60 mmol), aniline (5.64 g, 60 ACKNOWLEDGMENTS
mmol), and paraformaldehyde (3.61 g, 120 mmol) were mixed M. Baqar acknowledges the Ministry of Higher Education and
together and refluxed for 48 h in 1,4 dioxane (65 mL). The product Scientific Research of Libya and Azzaytuna University-Libya for
was concentrated using a vacuum evaporator and redissolved in ethyl a scholarship.


acetate followed by base-washed then once with water. The product
was then dried over sodium sulfate anhydrous, followed by vacuum
evaporation to afford a viscous product (yield: 9.12 g, 63%). REFERENCES
IR spectra (KBr, cm−1): 3325 (stretching of OH of methylol), 1500 (1) Ishida, H.; Agag, T., Eds. Handbook of Benzoxazine Resins;
(stretching of trisubstituted benzene ring), 1242 (asymmetric Elsevier: Amsterdam, 2011.

8124 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125


Macromolecules Article

(2) Ghosh, N. N.; Kiskan, B.; Yagci, Y. Prog. Polym. Sci. 2007, 32, (22) (a) Nair, C. P. R.; Bindu, R. L.; Ninan, K. N. J. Appl. Polym. Sci.
1344−1391. 2001, 81, 3371−3377. (b) Zhou, D. P.; Du, S.; Yu, L.; Liu, Z. J. Appl.
(3) (a) Sawaryn, C.; Landfester, K.; Taden, A. Macromolecules 2011, Polym. Sci. 2011, 121, 1938−1945.
44, 7668−7674. (b) Su, Y. C.; Kuo, S. W.; Yei, D. R.; Xu, H.; Chang, F. (23) Kumar, K. S. S.; Nair, C. P. R.; Ninan, K. N. Thermochim. Acta
C. Polymer 2003, 44, 2187−2191. (c) Kim, H. D.; Ishida, H. 2006, 441, 150−155.
Macromolecules 2003, 36, 8320−8329. (d) Kuo, S. W.; Wu, Y. C.;
Wang, C. F.; Jeong, K. U. J. Phys. Chem. C 2009, 113, 20666−20673.
(4) (a) Wang, Y. X.; Ishida, H. Polymer 1999, 40, 4563−4570.
(b) Chutayothin, P.; Ishida, H. Macromolecules 2010, 43, 4562−4572.
(c) Liu, C.; Shen, D.; Sebastian, R. M.; Marquet, J.; Schönfeld, R.
Macromolecules 2011, 44, 4616−4622.
(5) (a) Riess, G., Schwob, M., Guth, G., Roche, M., Lande, B. In:
Culbertson, B. M., McGrath, editors. Advances in Polymer Synthesis;
Plenum: New York, 1985. (b) Dunkers, J.; Ishida, H. J. Polym. Sci., Part
A: Polym. Chem. 1999, 37, 1913−1921. (c) Sudo, A.; Kudoh, R.;
Nakayama, H.; Arima, K.; Endo, T. Macromolecules 2008, 41, 9030−
9034.
(6) (a) Ishida, H.; Rodriguez, Y. Polymer 1995, 36, 3151−3158.
(b) Santhosh, K. S.; Reghunadhan, C. P.; Ninan, K. N. Thermochim.
Acta 2006, 441, 150−155.
(7) Wang, Y. X.; Ishida, H. J. Appl. Polym. Sci. 2002, 84, 1107−1113.
(8) Andreu, R.; Reina, J. A.; Ronda, J. C. J. Polym. Sci., Part A: Polym.
Chem. 2008, 46, 3353−3366.
(9) (a) McDonagh, A. F.; Smith, H. E. J. Org. Chem. 1968, 33, 8−12.
(b) McDonagh, A. F.; Smith, H. E. J. Org. Chem. 1968, 33, 1−8.
(10) (a) Dunkers, J.; Ishida, H. J. Polym. Sci., Part A: Polym. Chem.
1999, 37, 1913−1921. (b) Andreu, R.; Reina, J. A.; Ronda, J. C. J.
Polym. Sci., Part A: Polym. Chem. 2008, 46, 6091−6101. (c) Wang, Y.
X.; Ishida, H. Macromolecules 2000, 33, 2839−2847.
(11) (a) Wirasate, S.; Dhumrongvaraporn, S.; Allen, D. J.; Ishida, H.
J. Appl. Polym. Sci. 1998, 70, 1299−1306. (b) Schnell, I.; Brown, S. P.;
Low, H. Y.; Ishida, H.; Spiess, H. W. J. Am. Chem. Soc. 1998, 120,
11784−11795. (c) Kim, W. K.; Mattice, W. L. Comp. Theor. Polym. Sci.
1998, 8, 339−351. (d) Goward, G. R.; Shnell, I.; Brown, S. P.; Spiess,
H. W.; Kim, H. D.; Ishida, H. Magn. Reson. Chem. 2001, 39, S5−S17.
(e) Kim, H. D.; Ishida, H. J. Phys. Chem. A 2002, 106, 3271−3280.
(f) Goward, G. R.; Sebastiani, D.; Schnell, I.; Spiess, H. W.; Kim, H.
D.; Ishida, H. J. Am. Chem. Soc. 2003, 125, 5792−5800. (g) Kim, H.
D.; Ishida, H. Macromol. Symp. 2003, 195, 123−140. (h) Wang, C. F.;
Wang, Y. T.; Tung, P. H.; Kuo, S. W.; Lin, C. H.; Sheen, Y. C.; Chang,
F. C. Langmuir 2006, 22, 8289−8292. (i) Kuo, S. W.; Wu, Y. C.;
Wang, C. F.; Jeong, K. U. J. Phys. Chem. C 2009, 113, 20666−20673.
(12) (a) Laobuthee, A.; Chirachanchai, S.; Ishida, H.; Tashiro, K. J.
Am. Chem. Soc. 2001, 123, 9947−9955. (b) Chirachanchai, S.;
Laobuthee, A.; Phongtamrug, S. J. Heterocycl. Chem. 2009, 46, 714−
721.
(13) Kim, H. D.; Ishida, H. J. Appl. Polym. Sci. 2001, 79, 1207−1219.
(14) (a) Kiskan, B.; Yagci, Y.; Ishida, H. J. Polym. Sci., Part A: Polym.
Chem. 2008, 46, 414−420. (b) Kiskan, B.; Koz, B.; Yagci, Y. J. Polym.
Sci., Part A: Polym. Chem. 2009, 47, 6955−6961.
(15) Kudoh, R.; Sudo, A.; Endo, T. Macromolecules 2009, 42, 2327−
2329.
(16) Sudo, A.; Du, L. C.; Hirayama, S.; Endo, T. J. Polym. Sci., Part A:
Polym. Chem. 2010, 48, 2777−2782.
(17) Oie, H.; Sudo, A.; Endo, T. J. Polym. Sci., Part A: Polym. Chem.
2010, 48, 5357−5363.
(18) (a) Baqar, M.; Agag, T.; Ishida, H.; Qutubuddin, T. J. Polym. Sci.,
Part A: Polym. Chem. 2012, 50, 2275−2285. (b) Baqar, M.; Agag, T.;
Ishida, H.; Qutubuddin, S. Polymer 2011, 52, 307−317. (c) Baqar, M.;
Agag, T.; Ishida, H.; Qutubuddin, S. React. Funct. Polym. 2012, http://
dx.doi.org/10.1016/j.reactfunctpolym.2012.04.017.
(19) Dunkers, J.; Ishida, H. Spectrochim. Acta 1995, 51A, 1061−1074.
(20) (a) Agag, T.; Takeichi, T. Macromolecules 2003, 36, 6010−6017.
(b) Dunkers, J.; Ishida, H. Spectrochim. Acta 1995, 51A, 855−867.
(21) (a) Holopainen, T.; Alvila, L.; Rainio, J.; Pakkanen, T. T. J. Appl.
Polym. Sci. 1998, 69, 2175−2185. (b) Holopainen, H.; Alvila, L.;
Pakkanen, T. T.; Rainio, J. J. Appl. Polym. Sci. 2003, 89, 3582−3586.

8125 dx.doi.org/10.1021/ma301963d | Macromolecules 2012, 45, 8119−8125

You might also like