You are on page 1of 6

Thermochimica Acta 595 (2014) 22–27

Contents lists available at ScienceDirect

Thermochimica Acta
journal homepage: www.elsevier.com/locate/tca

The curing kinetics and thermal properties of epoxy resins cured by


aromatic diamine with hetero-cyclic side chain structure
Xuhai Xiong a , Rong Ren a , Siyang Liu a , Shaowei Lu a , Ping Chen a,b, *
a
Liaoning Key Laboratory of Advanced Polymer Matrix Composites Manufacturing Technology, Shenyang Aerospace University, Shenyang 110136, China
b
State Key Laboratory of Fine Chemicals, School of Chemical Engineering, Dalian University of Technology, Dalian 116012, China

A R T I C L E I N F O A B S T R A C T

Article history: The use of novel aromatic diamine containing phthalide structure (BAPP) as curing agent of
Received 18 May 2014 diglycidylether of bisphenol A (DGEBA) epoxy resin has been studied. BAPP can react with epoxy
Received in revised form 28 July 2014 monomers by an epoxy-amine condensation mechanism, which renders phthalide cardo incorporate
Accepted 29 July 2014
into the network structure of the thermoset. The curing behavior and kinetics have been investigated by
Available online 1 August 2014
differential scanning calorimetry (DSC) and the kinetic analysis of non-isothermal cure shows that
autocatalytic model is suitable to describe the cure mechanism. The thermal–mechanical properties have
Keywords:
been determined with dynamic mechanical analysis (DMA) and the activation energies for relaxation
Aromatic amine
Phthalide structure
were calculated according to the Arrhenius law. Thermogravimetric analysis exhibits a lower initial
Epoxy resin decomposition temperature, decreased rate of decomposition and higher char yield compared with
Curing kinetics DGEBA cured with commercially available 4,40 -diaminodiphenylsulfone (DDS).
Thermal–mechanical properties ã 2014 Published by Elsevier B.V.

1. Introduction increasing molecular chain of curing agent will reduce crosslinked


density and further improve the toughness of network, but heat
Curing agents play an important role in determining the practical resistance is decreased [6]. Therefore, the chain-extended aromatic
use of epoxy resin [1–3]. The excellent properties, such as mechanical amines with multiaryl skeleton may provide a good balance
strength, heat resistance, durability and adhesiveness, are obtained between mechanical and thermal properties.
by reacting the linear epoxy resin with suitable curatives to form Phthalide moieties are the ever-popular hetero-cyclic side
three-dimensional crosslinked thermoset networks [4,5]. Therefore, cardo groups and incorporated into polymer main chains to
design and development of novel curing agent is one of effective improve solubility, heat resistance and thermal oxidative stability
methods to improve properties of epoxy resin [2,5–8]. [11–13]. Phenolphthalein and its derivations have been used
Despite the fact that epoxy resins could be cured by many types of largely as aromatic bisphenol monomers for preparing high-
hardeners, aromatic amines retain a prominent position in high-tech performance thermoplastics and thermosetting resins. According-
fields [9,10]. Epoxy resins cured with aromatic amine generally ly, phthalide-containing epoxy compounds were designed and
provide enhanced environmental (hydrolytic) stability, outstanding synthesized, and various epoxy resin systems modified by
heat-resistant and mechanical properties [5,9]. 4,40 -diaminodi- polymers with phthalide cardo structure were developed
phenyl methane (DDM), m-phenylene diamine (m-PDA) and [14–16]. Although phthalide-modified epoxy systems show a lot
4,40 -diaminodiphenylsulfone (DDS) are principal commercial of advantages, a more cost-effective approach to introduce
aromatic-amine curing agents. These traditional aromatic amines phthalide moiety into epoxy networks using phthalide-contained
with low molecular weight, however, bring on highly crosslinked aromatic amine curing agents, which are of large molecular weight
materials, which are seized of greater brittleness. Generally, and have the potential to overcome the defects of traditional
aromatic amines, has been not paid more attention.
The present work aims to investigate the curing kinetics and
* Corresponding author at: Liaoning Key Laboratory of Advanced Polymer Matrix mechanism of diglycidylether of bisphenol A (DGEBA) epoxy resin
Composites Manufacturing Technology, Shenyang Aerospace University, No. 37, cured with aromatic diamine containing phthalide structure
Daoyi South Avenue, Daoyi Development District, Shenyang 110136, China. Tel.: +86
24 89723970.
(BAPP) by using the non-isothermal DSC. The thermal properties
E-mail addresses: xiongxuhai@126.com (X. Xiong), chenping_898@126.com including thermal mechanical and thermal stability of cured
(P. Chen). material will be also studied by means of DMA and TGA.

http://dx.doi.org/10.1016/j.tca.2014.07.027
0040-6031/ ã 2014 Published by Elsevier B.V.
X. Xiong et al. / Thermochimica Acta 595 (2014) 22–27 23

2. Experimental

2.1. Materials

Commercially available diglycidylether of bisphenol A


(DGEBA)-based epoxy resin, with an epoxy equivalent of about
185–210 and average equivalent weight of 196, was supplied by
Wuxi Resin Works, China, and dried at 100  C in vacuum for 1 h
before use. The hardener, 3,3-bis[4-(4-aminophenoxy)phenyl]
phthalide (BAPP), was synthesized according to the reported
method [17]. The molecular structures of chemical structures
DGEBA and BAPP are illustrated in Scheme 1.

2.2. Preparation of DGEBA/BAPP blend

The prepolymer DGEBA/BAPP was prepared by mixing


stoichiometric DGEBA and BAPP at 90  C under vigorous mechani-
cal stirring for about 5 min, and then a homogeneous, visually
transparent mixture was obtained. The mixture was poured Fig. 1. Non-isothermal DSC thermographs of DGEBA/BAPP reactions.
directly into preheated teflon mold and then the mold was
transferred into an oven at 100  C under vacuum for 5 min to drive 3. Results and discussion
off entrapped bubbles. The following temperature programs were
used to finish the cure process in an air convection oven at 120  C 3.1. DSC analysis
for 1 h, 160  C for 2 h plus a post-cure period at 180  C for 4 h.
Finally, the casting was removed from the mold and characterized. 3.1.1. Reactivity of DGEBA/BAPP
The non-isothermal curing reactions of DGEBA/BAPP were
studied using DSC at a heating rate of 5, 7.5, 10 and 15  C/min and
2.3. Characterization the dynamic DSC thermograms and characteristic parameters are
shown in Fig. 1 and Table 1, respectively. It can be seen that there is
Differential scanning calorimetry (DSC) measurements were a single exothermic peak in each DSC curves and heating rate has a
conducted with a PerkinElmer Diamond DSC instrument. The great influence on the shape of the exothermic curves. The
DGEBA/BAPP mixture (about 7–9 mg mass) was loaded into sealed exothermic peak is shifted to higher temperature and peak width is
aluminum pan. All DSC experiments were performed under N2 extended with increasing heating rate. In addition, the total heat
protection and run twice, the first scan was conducted from 25 to release of cure reaction is independent of heating rate. In order to
300  C at different heating rates of 5, 7.5, 10 or 15  C/min and the determine the glass transition temperature (Tg) of DGEBA/BAPP,
second scan was conducted from 25 to 200  C at a heating rates of the cured samples were carried out the second DSC test at a
20  C/min. heating rate of 20  C/min. As seen from Fig. 2, Tg value shows a
Dynamic mechanical analysis (DMA) was done on a TA strong dependence on the heating rate of the first DSC test and
Instruments Q800 DMA with an amplitude of 20 mm and a decreases by about 10  C with shifting of heating rate from 5 to
temperature ramp rate of 3  C/min. The scanning temperature was 15  C/min (see Table 1). The believable cause is that the increasing
from 25 to 200  C and the driving frequencies were 1.0, 2.0, 5.0, 10 heating rates reduce the network chains packing density, resulting
and 20 Hz. The cured DGEBA/BAPP specimens were cut to in a decrease of steric hindrance to molecular motion.
dimensions of 30 mm  6 mm 1 mm for the tension mode.
Thermogravimetric analysis (TGA) was performed on a Perki- 3.1.2. Activation energy of the non-isothermal cure
nElmer TGA-7 thermal analyzer and the cured DGEBA/BAPP sample The curing of epoxy resins is a multi-step chemical process that
(around 5 mg) was heated from 25 to 700  C at a heating rate of is complicated by the physical processes of gelation and vitrifica-
20  C/min under the purified nitrogen flow rate of 60 mL/min. tion [9,10]. Fortunately, the isoconversional analysis is a powerful

Scheme 1. Chemical structure of DGEBA and BAPP.


24 X. Xiong et al. / Thermochimica Acta 595 (2014) 22–27

Table 1
DSC analysis of DGEBA/BAPP resin at different heating rates.

b (K/min) Tonseta ( C) Tpeakb ( C) Tendc ( C) 4Hd (J/g) Tge ( C)


5 131.7 170.5 221.3 212.6 157.1
7.5 133.8 178.6 230.2 221.4 153.4
10 136.1 185.8 238.9 217.5 151.2
15 142.2 196.1 260.4 228.4 147.8
a
The onset cure temperatures.
b
The peak exothermic temperature.
c
The end cure temperatures.
d
Total heat of cure reaction.
e
Glass transition temperature measured by DSC.

tool to explore the evolvement of cure mechanism by monitoring


the variations in the effective activation energy with the extent of
curing [18–21]. To perform isoconversional analysis, the original
DSC data on exothermic peak were transformed into the fractional
conversion (a) versus temperature curves at various heating rates,
and the corresponding plots are shown in Fig. 3. The a values were Fig. 3. Conversion degree as a function of temperature.
obtained by the integration of the exothermic peak based on the
following equation: suggests that the process follows a single step kinetic model. After
a ¼ Ha =Htotal (1)
a > 60%, the values of Ea decrease gradually to as low as 50 kJ as
the degree of conversion increases. The same phenomenon was
where Ha is the fractional enthalpy and Htotal is the total enthalpy observed in the non-isothermal or isothermal cure of DGEBA/1,3-
of the cure reaction. phenylenediamine (m-PDA) using Vyazovkin method [19]. The
It is clear from Fig. 3 that all a values increase very slowly at the decrease in Ea may indicate a change of the reaction mechanism.
beginning of curing and when the samples were heated to given However, decreased amplitude of the Ea value is smaller than that
temperatures, the a values show a sharply increase and then level reported in the other works [18,19]. The reason may be that when
off. Moreover, to obtain the same a value, the required temperature a > 60% epoxy curing system in the present consideration has
is increased with increasing heating rate. passed the gel point and loses the ability to flow. In other hand,
A modified integral method proposed by M.J. Starink based on from Figs. 2 and 3 it is known the actual experimental temperature
Kissinger–Akahira–Sunose method, which is of more accurate, was (a > 60%) is always above the glass transition temperature, hereat,
used to estimates the activation energies (Ea) for different the mixture of polymer network and monomer is not frozen and
conversion levels (a). Starink equation is expressed as follows [22]: capable of engaging in further polymerization. Based on the
aforementioned facts, it is reasonable to expect that the decrease in
lnðb=T 1:92
a Þ ¼ const  1:0008  ðEa =RT a Þ (2)
Ea results from increasing diffusion control.
where b is the heating rate, R is the universal gas constant, Ta is the
absolute temperature at a fixed value of the variable a. 3.1.3. Model-fitting kinetic analysis
The activation energy can be calculated from the slope of the Reaction rate equation for thermosetting system is generally
linear plot of ln(b/Ta1.92) against 1/Ta. The resulting Ea-dependen- expressed
cies are shown in Fig. 4. It can be seen that the Ea values are around
da Ea
47.5–60.9 kJ/mol, the average value is 56.1 kJ/mol, which is quite ¼ Aexpð Þf ðaÞ (3)
dt RT
typical for epoxy/aromatic amine copolymerizations [6,19]. In the
interval 0.05  a  0.6, the Ea values are practically constant which

Fig. 4. Variation of effective activation energy calculated based on Starink method


Fig. 2. Tgs of cured DGEBA/BAPP measured by DSC at a heating rate of 20 K/min. with conversion.
X. Xiong et al. / Thermochimica Acta 595 (2014) 22–27 25

Fig. 5. Plots of normalized y(a) and z(a) against a.


Fig. 6. Plots of ln[y(a)] versus ln[ap(1  a)].

where f(a) is the reaction model. Curing reaction of epoxy/amine


examination shows that the values of am, ap1 and ap (the
system is assumed to follow autocatalytic reaction model (two-
conversion degree on the maximum of curing exothermic peak)
parameter SB model) [5,23,24]:
satisfy the conditions: 0 < am < ap < ap1 and ap1 6¼ 0.632, which
da Ea strongly indicate that the reactions of DGEBA/BAPP follow
¼ Aexpð Þam ð1  aÞn (4)
dt RT autocatalytic reaction model.
where f ðaÞ ¼ am ð1  aÞn , m and n are reaction orders. When m = 0, The kinetic exponents n and m and the pre-exponential factor A
Eq. (4) is transformed into n-order model. were obtained according to Eq. (8). The parameters, p, were first
Kinetic parameters (m, n) of the cure reaction were estimated determined from the values of am and then a series of the plots of
using the Málek method [24–26]. In this method two parameters ln[y(a)] with respect to ln[ap(1  a)] at different heating rates
am and ap1, which are the maximum of y(a) and z(a), respectively, were constructed and shown in Fig. 6. These data show good
are first obtained based on Eqs. (5) and (6). linearity. The values of n, m and ln A were calculated from the slope
and intercept of fitting straight lines. The constant kinetic
da parameters thus were determined and listed in Table 2. Then,
yðaÞ ¼ ð ÞexpðxÞ (5)
dt the averaged m, n and ln A along with previously calculated Ea are
introduced into Eq. (4), and eventually the explicit reaction rate
equation can be obtained as Eq. (9):
da T
zðaÞ ¼ pðxÞð Þ (6)
dt b da 56100 0:232
¼ 8:86  105 expð Þa ð1  aÞ1:265 (9)
where x is the reduced activation energy, x ¼ Ea =RT; pðxÞ was dt RT
calculated using the numerical approach of Senum–Yang Eq. (7): In Fig. 7, a comparison between experimental and theoretical
values is presented to check the predictability of Eq. (9). Obviously,
x3 þ 18x2 þ 88x þ 96 the predicted rate is in fair accordance with the experimental rate.
pðxÞ ¼ (7)
x4 þ 20x3 þ 120x2 þ 240x þ 120 This demonstrates the autocatalytic reaction model is suitable to
By combining Eqs. (4) and (5) and yields Eq. (8)

ln½yðaÞ ¼ lnA þ nln½ap ð1  aÞ (8)


where p ¼ m=n ¼ am =ð1  am Þ
The average Ea value calculated by the Starink method was
introduced into Eqs. (5) and (6) and and function curves of y(a),
which were normalized in order to simplify, and z(a) could be
constructed, as shown in Fig. 5. This figure clearly indicates that
conversion am associated with the peak values of y(a) curves are
located in 0.10–0.18, and the z(a) curves show a “C” shaped contour
experiencing a practically isoconversional peak value conversion,
ap1, within 0.50–0.54. The values am and ap1 corresponding to
different heating rates are shown in Table 2. A more detailed

Table 2
Characteristic peak conversion values and calculated kinetic parameters for SB
model at different heating rates.

b (K/min) ap am ap1 p m n ln A
5 0.508 0.159 0.535 0.189 0.260 1.202 13.64
7.5 0.479 0.123 0.516 0.140 0.216 1.271 13.76
10 0.484 0.155 0.516 0.183 0.181 1.162 13.57
15 0.471 0.136 0.508 0.157 0.273 1.424 13.81
Mean 0.486 0.143 0.519 0.168 0.232 1.265 13.70 Fig. 7. Comparison of predicted autocatalytic reaction model with experimental
data.
26 X. Xiong et al. / Thermochimica Acta 595 (2014) 22–27

Fig. 10. TGA thermograms of cured DGEBA/BAPP systems.


Fig. 8. DMA thermograms of cured DGEBA/BAPP systems.

describe the progress of the non-isothermal curing reaction of


DGEBA/BAPP system. dln f
DE ¼ R (10)
dð1=T g Þ
3.2. Dynamic mechanical thermal analysis
where 4E is the activation energy for specific relaxation, Tg denotes
the corresponding relaxation temperature, R is the universal gas
The DMA curves conducted at a driving frequency of 1.0, 2.0, 5.0,
constant, and f is the loading vibration frequency.
10 and 20 Hz for the DGEBA/BAPP networks are demonstrated in
A series of f and Tg was obtained from the multi-frequency DMA
Fig. 8. Generally, peak temperature for tan d or loss modulus is
measurements under fixed heating rate and the typical kinetic
identified as the glass transition temperature (Tg) and the area of
plots of ln f versus 1/Tg were drawn. According to the plots shown
peak is associated with the amount of energy loss caused by
in Fig. 9, ln f against 1/Tg presents a good linear correlation and 4E
molecular motion. Seen from Fig. 8, tan d curves show single peak
was calculated from the value of the slope. As a result, the
at a temperature range from 140 to 200  C, which is an indication
activation energy values calculated based on loss modulus peak
that DGEBA/BAPP networks have homogeneous microstructure.
temperature and tan d peak temperature are 638.8 and
Besides, it is obvious that storage modulus, loss modulus and
587.9 kJ/mol, respectively. The results are higher than the 4E
relaxation peak express strong frequency dependency. As loading
values (563 and 538 kJ/mol) for glass transition of cured DGEBA/
frequency increases, the network becomes somewhat stiffer. This
DDM system estimated by Raditoiu et al. [27]. This may be ascribed
is because the motion of network segments can follow the change
to more rigid molecular segment of BAPP than that of DDM.
of driving force when smaller frequency is employed. The
increasing frequency also leads to a shift of loss modulus peak
3.3. Thermogravimetric analysis
and tan d peak to a higher temperature and the shift of two peaks
shows strong regularity. Herein, the Arrhenius law is adopted to
The thermal stability of the cured DGEBA/BAPP was assessed
correlate this interrelationship for calculating the activation
with TGA and compared with that of DGEBA cured with 4,40 -
energy of relaxation [26]. The activation energies of relaxation
diaminodiphenylsulfone (DDS), which was obtained from the
were determined based on Eq. (10).
literature [16]. Fig. 10 compares the thermal degradation behavior
of DGEBA/BAPP and DGEBA/DDS in nitrogen atmosphere. DGEBA/
BAPP system exhibits a lower temperature of initial decomposition
(Ti), slower rate of weight loss and higher char yield. The decreased
Ti may be attributed to the lower crosslinking density and the
relatively weaker tertiary amine crosslinks because of stronger
basicity of BAPP than that of DDS [28]. On the other hand, the
present of more aromatic structure may be the most significant
factor contributing to the lower rate of weight loss and higher char
yield. The latter indicates that the DGEBA/BAPP system is of higher
limiting oxygen index (LOI) and better flame-retardant property
according to the method of van Krevelen and Nijenhuis [29].

4. Conclusions

A novel epoxy system was developed using aromatic amine


containing phthalide cardo structure as curing agent. The curing
behavior and kinetics were studied in detail by non-isothermal
DSC. The average values of the apparent activation energy obtained
by Starink equation were 56.1 kJ/mol. Model-fitting kinetic analysis
showed the
curing mechanism of DGEBA/BAPP system followed autocata-
Fig. 9. Plots of ln f versus 1/Tg. lytic model. The viscoelastic properties of the cured material were
X. Xiong et al. / Thermochimica Acta 595 (2014) 22–27 27

characterized by DMA and single peak appeared in tan d curve at a [13] M.S. Lin, E.M. Pearce, Polymers with improved flammability characteristics. I.
temperature range from 140 to 200  C. Activation energies for the Phenolphthalein-related homopolymers, J. Polym. Sci. Part A: Polym. Chem. 19
(1981) 2659–2670.
glass transition determined from tan d and loss modulus peaks are [14] B. Everatt, A.H. Haines, B.P. Stark, Investigation by carbon-13 and proton NMR
638.8 and 587.9 kJ/mol, respectively. The results obtained from spectroscopy of the diglycidyl derivative of phenolphthalein, Angew.
TGA analysis indicated the cured material was provided with a Makromol. Chem. 56 (1976) 157–162.
[15] J.Y. Li, P. Chen, Z.M. Ma, Reaction kinetics and thermal properties of cyanate
decreased rate of thermal decomposition and a higher char yield. ester-cured epoxy resin with phenolphthalein poly(ether ketone), J. Appl.
Polym. Sci. 111 (2009) 2590–2596.
[16] X.H. Xiong, P. Chen, J.X. Zhang, Preparation and properties of high performance
Acknowledgements phthalide-containing bismaleimide modified epoxy matrices, J. Appl. Polym.
Sci. 121 (2011) 3122–3130.
The financial supportfrom NationalNatural Science Foundation of [17] X.H. Xiong, P. Chen, Q. Yu, Synthesis and properties of chain-extended
bismaleimide resins containing phthalide cardo structure, Polym Int. 59 (2010)
China (No. 51303107), National Defense 12th Five-Year Fundamental 1665–1672.
Research Program (No. A352*******) and program for Liaoning [18] S. Vyazovkin, N. Sbirrazzuoli, Mechanism and kinetics of epoxy-amine cure
Excellent Talents in University (No. LR2013002) is gratefully studied by differential scanning calorimetry, Macromolecules 29 (1996)
1867–1873.
acknowledged. [19] N. Sbirrazzuoli, S. Vyazovkin, A. Mititelu, A Study of epoxy-amine cure kinetics
by combining isoconversional analysis with temperature modulated DSC and
dynamic rheometry, Macromol. Chem. Phys. 204 (2003) 1815–1821.
References [20] S. Vyazovkin, N. Sbirrazzuoli, Isoconversional kinetic analysis of thermally
stimulated processes in polymers, Macromol. Rapid Commun. 2 (2006)
[1] A. Catalani, M.G. Bonicelli, Kinetics of the curing reaction of a diglycidyl ether 1515–1532.
of bisphenol A with a modified polyamine, Thermochim. Acta 438 (2005) [21] N. Sbirrazzuoli, A.M. Mija, L. Vincent, Isoconversional kinetic analysis of
126–129. stoichiometric and off-stoichiometric epoxy-amine cures, Thermochim. Acta
[2] T. Maity, B.C. Samanta, S. Dalai, A.K. Banthia, Curing study of epoxy resin by 447 (2006) 167–177.
new aromatic amine functional curing agents along with mechanical and [22] M.J. Starink, The determination of activation energy from linear heating rate
thermal evaluation, Mater. Sci. Eng. A 464 (2007) 38–46. experiments: a comparison of the accuracy of isoconversion methods,
[3] E.S. Balcerzaka, H. Janeczeka, B. Kaczmarczyka, Epoxy resin cured with Thermochim. Acta 404 (2003) 163–176.
diamine bearing azobenzene group, Polymer 45 (2004) 2483–2493. [23] C. Li, C.M. Zuo, H. Fan, Novel silicone aliphatic amine curing agent for epoxy
[4] D. Rosu, F. Mustata, C.N. Cascaval, Investigation of the curing reactions of some resin: 1,3-bis(2-aminoethylaminomethyl) tetramethyldisiloxane. 1. Non-iso-
multifunctional epoxy resins using differential scanning calorimetry, Ther- thermal cure and thermal decomposition property, Thermochim. Acta 545
mochim. Acta 370 (2001) 105–110. (2012) 75–81.
[5] M. Ghaemy, M. Barghamadi, H. Behmadi, Cure kinetics of epoxy resin and [24] Z.Y. Bu, J.J. Hu, B.G. Li, Novel silicon-modified phenolic novolac resins: non-
aromatic diamines, J. Appl. Polym. Sci. 94 (2004) 1049–1056. isothermal curing kinetics, and mechanical and thermal properties of their
[6] M.F. Mustafa, W.D. Cook, T.L. Schiller, Curing behavior and thermal properties biofiber-reinforced composites, Thermochim. Acta 575 (2014) 244–253.
of TGDDM copolymerized with a new pyridine-containing diamine and with [25] E.N. Karasinski, M.G.D. Luzb, C.M. Lepienski, Nanostructured coating based on
DDM or DDS, Thermochim. Acta 575 (2014) 21–28. epoxy/metal oxides: kinetic curing and mechanical properties, Thermochim.
[7] G.R. Saada, E.E.A. Elhamidb, S.A. Elmenyawyb, Dynamic cure kinetics and Acta 569 (2013) 167–176.
thermal degradation of brominated epoxy resin–organoclay based nano- [26] C. Li, H. Fan, J.J. Hu, Novel silicone aliphatic amine curing agent for epoxy
composites, Thermochim. Acta 524 (2011) 186–193. resin: 1,3-bis(2-aminoethylaminomethyl) tetramethyldisiloxane. 2. Isother-
[8] Y. Zhang, S. Vyazovkin, Macromol. Chem. Phys. 206 (2005) 342–348. mal cure, and dynamic mechanical property, Thermochim. Acta 549 (2012)
[9] T. Dyakonov, Y. Chen, K. Holland, Thermal analysis of some aromatic amine 132–139.
cured model epoxy resin systems—I: materials synthesis and characterization, [27] V. Raditoiu, R. Gabor, C.A. Nicolae, L. Dumitrache, A. Raditoiu, M. Mihailescud,
cure and post-cure, Polym. Degrad. Stab. 53 (1996) 211–242. Damping properties of some epoxy resin–liquid crystalline dyestuff compo-
[10] B.A. Rozenberg, Kinetics, thermodynamics and mechanism of reactions of sites, U.P.B. Sci. Bull. Series A 76 (2014) 223–230.
epoxy oligomers with amines, Adv. Polym. Sci. 75 (1985) 113–165. [28] T. Dyakonov, P.J. Mann, Y. Chen, Thermal analysis of some aromatic amine
[11] C.P. Yang, Y.Y. Su, J.M. Wang, Synthesis and properties of organosoluble cured model epoxy resin systems—II: residues of degradation, Polym. Degrad.
polynaphthalimides based on 1,4,5,8-naphthalene tetracarboxylic dianhy- Stab. 54 (1996) 67–83.
dride, 3,3-bis[4-(4-aminophenoxy)phenyl] phthalide, and various aromatic [29] D.W. van Krevelen, K. te Nijenhuis, Properties of Polymers, Elsevier, New York,
diamines, J. Polym. Sci. Part A: Polym. Chem. 44 (2006) 940–948. 2009, pp. 856.
[12] Z.G. Wang, T.L. Chen, J.P. Xu, Gas and water vapor transport through a series of
novel poly(arylether sulfone) membrane, Macromolecules 34 (2001)
9015–9022.

You might also like