Sei sulla pagina 1di 12

Arabian Journal of Chemistry (2017) xxx, xxx–xxx

King Saud University

Arabian Journal of Chemistry


www.ksu.edu.sa
www.sciencedirect.com

ORIGINAL ARTICLE

Curing and thermal degradation of diglycidyl ether


of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly
hardeners
N. Tudorachi, F. Mustata *

‘‘Petru Poni” Institute of Macromolecular Chemistry, Gr. Ghica Voda Alley, 41A, 700487 Iasi, Romania

Received 12 April 2017; accepted 12 July 2017

KEYWORDS Abstract The paper presents the curing of diglycidyl ether of bisphenol A (DGEBA) with two nat-
Natural hydroxy acids; ural hydroxy acids and the thermal behavior of the cured products. The kinetics of the curing reac-
Curing reaction; tions were evaluated from the differential scanning calorimetry (DSC) curves using Ozawa,
Thermosetting resin; Kissinger methods and ‘‘Thermokinetics-3” software (Netzsch). Thermal decomposition behavior
Thermal properties of the cured products up to 600 °C (in nitrogen) was made using simultaneous TG/FT-IR/MS tech-
nique. The kinetic parameters were obtained using the same software which shows that the thermal
degradation takes place in three or four consecutive steps. The chemical composition of gases
evolved at the thermal decomposition has been identified using FT-IR and MS techniques. The
main products are: water, carbon dioxide, acids, anhydrides, aliphatic and aromatic hydrocarbons.
Ó 2017 Production and hosting by Elsevier B.V. on behalf of King Saud University. This is an open access
article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction the fields of the application for these resins are very wide, from
industrial applications (aerospace, car or ship industry, electri-
Epoxy resins are some of the most important products inten- cal and electronic devices, etc.) to the fields of household and
sively used as thermosetting materials. Due to the diversity medical items (Ellis, 1993; Lee and Neville, 1972; May,
of their chemical compositions and the good properties 1988). However, as all the synthetic or natural polymer mate-
obtained after crosslinking (impact resistance, high electrical rials exhibit a number of properties that limit their use. One of
insulation and resistance to chemical substances and moisture) the main deficiencies is that these materials are not environ-
mentally friendly. To combat this deficiency, epoxy resins
* Corresponding author. and curing agents obtained from natural resources were used
E-mail address: fmustata@icmpp.ro (F. Mustata). as potential biodegradable promoters in the formulation of
Peer review under responsibility of King Saud University. epoxy resins composition (Alam et al., 2014; Jaillet et al.,
2013; Li et al., 2006; Miyagawa et al., 2004; Mustata et al.,
2014, 2013a, 2011; Mustata and Tudorachi, 2010; Park
et al., 2004; Roudsari et al., 2014; Das and Karak, 2009; Liu
Production and hosting by Elsevier
et al., 2012). The use of crosslinking agents obtained from
http://dx.doi.org/10.1016/j.arabjc.2017.07.008
1878-5352 Ó 2017 Production and hosting by Elsevier B.V. on behalf of King Saud University.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
2 N. Tudorachi, F. Mustata

the natural resources not always give crosslinking products heating rates under nitrogen atmosphere. The majority of the
with higher mechanical and thermal properties than those sample was placed in aluminum moulds with the next dimen-
obtained with non-environmentally friendly curing agents, sions: length = 3  102 m, width = 2  102 m height of about
but are sometimes advantageous because they are nontoxic 0.5  102 m and cured in an oven, after next thermal cycle: at
and come from renewable sources (Jaillet et al., 2013; Li 80 °C (1 h), 120 °C (1 h) and post cured at 160 °C for 5 h. The
et al., 2006). Most of the thermosetting materials are based fact that DSC curves of the cured samples present only
on epoxy resin as matrix, mixed mainly with hardeners, curing exothermic peak suggest that the curing reaction was complete.
catalysts, reactive or inert diluents, fillers, dyes, etc. Catalysts Finally, the cured samples were grounded as fine powder and
have a great influence on the processing conditions and the used for thermal analysis.
final characteristics of the crosslinked products. By way that
is prepared the thermosetting material, the catalysts are kept 2.3. Measurements
in the resin matrix. In our case, triethylbenzylammonium chlo-
ride (TEBAC) was used as the catalyst, and is embedded in the The epoxy equivalent was determined as in literature by titra-
final product. Considering this research direction, we try to use tion with hydrogen bromide (ASTM D 1652-04) and has 190
as hardening agents for epoxy resins two hydroxy acids of nat- value. Fourier transform infrared (FT-IR) spectra were
ural origin (citric and tartaric acids). Citric acid derivatives recorded with a Vertex 70 spectrophotometer (Bruker-
have also been used as crosslinkers, but they contained in their Germany) on KBr disks, in the range of 4000–400 cm–1. The
chemical structure large amounts of aliphatic or aromatic ami- penetration resistance of the crosslinked products was
nes (Galego et al., 1992, 1996). These two acids contain in their obtained using the Shore hardness test and recorded with a
chemical structure two or three carboxyl groups which can Shore D-type apparatus (SAUTER HB-Germany). The depth
react with epoxy groups. On the other hand, besides the car- penetration of the device head in tested sample as the average
boxyl groups these acids contain one or two hydroxyl groups value of 5 determinations, was calculate as Shore hardness.
which in conditions of higher temperatures can react with Vickers hardness was obtained with Vickers hardness tester
epoxy groups leading to a product with high crosslinking (Shimadzu Japan) under load of 1.961 N and dwell time of
degree. 12 s. The values of hardness were reported as an average of
The aim of this paper is the study of DGEBA curing of with 6 determinations. Degree of cross-linking of epoxy resins was
citric and tartaric acids and thermal characterization of the measured based on the literature method by extraction into
cured products using the simultaneous analysis TG/FT-IR/ solvent mixture (xylene/DMF, 1:1 v/v) [Patil et al., 2017].
MS. The cured samples, ground as fine grain, placed in a stainless
steel wire cloth, was immersed into solvent mixture, main-
2. Experimental tained at room temperature (25 °C) one day, filtered and dried
under vacuum at 150 °C to constant mass. The gel content
2.1. Materials (Gc) was calculated with the next equation:
Gc ¼ ðWi=WfÞ  100 ð1Þ
Citric acid (CA) (m.p. = 152–154 °C), L-tartaric acid (TA) (m.
where: Wi is the initial weight and Wf is the weight after 24 h
p. = 170–172 °C) and TEBAC (m.p. = 190–193 °C) were ana-
of extraction.
lytical grade products and used as received. The epoxy resin of
The SEM/ESEM – EDAX – QUANTA 200 apparatus was
DGEBA type provided by Sintofarm, Bucharest, was a com-
used to register the scanning electron microscopy images. The
mercial product with a molecular weight of about 390 g mol1,
next parameters were used: high vacuum, field emission fila-
epoxy equivalent of 190, viscosity of 15,000 mPas (measured
ment 15 kV accelerating voltage and magnification of 5000.
with Brookfield Viscometer, stil 4, at 25 °C) and used without
The samples obtained by fracturing after cooling with liquid
further purification.
nitrogen and coated with fine gold layer were used for
registration.
2.2. The sample preparation for the curing test and thermal The DSC measurements for the curing reactions were con-
experiment ducted on a DSC 200 F3 Maia device (Netzsch, Germany) in a
dynamic mode. The DSC curves were registered at different
The test samples entitled DGEBA/CA and DGEBA/TA were heating rates (5, 10, 15 °C min1) under nitrogen atmosphere
obtained by manually mixing until to complete homogeniza- (50 mL min1) with the temperature ranging from 25 to
tion of the liquid DGEBA resin, CA and TA as fine powder, 300 °C. Samples weighing up to 12 mg, were weighed carefully,
at molar ratio 1/0.75 epoxy ring/carboxylic proton in presence encapsulated in sealed aluminum pans and heated in the pres-
of TEBAC as catalyst (1% based on the monomers weight). ence of an empty crucible as reference, for each heating rate. It
For comparison, samples obtained in the absence of catalyst was recorded heat flow versus temperature. Previously, the
were also tested. The choice of the 1/0.75 epoxy ring/carboxyl device was calibrated with indium, under nitrogen atmosphere,
proton ratio is justified by the presence of hydroxyl groups in for all the three heating rates. The measurements were made
the structure of two acids, which may exhibit etherification consecutive from the same sample in the same day
reactions with the epoxide groups excess, with the increase of (Vyazovkin et al., 2014).
the crosslinking temperature. Then, for the removal of the The kinetic parameters of the curing reactions were deter-
included air, the samples were placed in an oven at 40 °C mined using model-freeestimation methods of Kissinger (KS)
and degassed under vacuum in about 30 min. After cooling and Ozawa-Flynn-Wall (OFW) (Kissinger, 1957; Ozawa,
at 0 °C, a small part was used for DSC measurement at three 1976; Flynn and Wall, 1966), based on the next Eqs. (2) and (3):

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
Curing and thermal degradation of diglycidyl ether 3

lnðb=T2p Þ ¼ Ek =RTp  lnðAR=Ek Þ ðKS equationÞ ð2Þ made of polytetrafluorethylene with 1.5 mm diameter and 1 m
length, maintained at 190 °C to TGA-IR external modulus
and equipped with MCT (Mercury Cadmium Telluride) detector.
ln b ¼ C  1:052ðEo =RTp Þ ðOFW equationÞ ð3Þ FT-IR spectra were registered in 3D size with OPUS 6.5 soft-
ware. For characterization of the gases by mass spectroscopy,
where: A is the pre-exponential factor, b is the heating rate, C
the evolved gases are transferred at the mass spectrometer
is a constant, Ek and Eo are the activation energies obtained
through a quartz capillary with 2 m length and 75 lm diame-
with KS and OFW equations, Tp is the maximum exothermic
ter, maintained at 290 °C. The working parameters were: ion-
peak temperature and R is the gas constant.
izing energy with electron impact of 70 eV, vacuum 105 mbar,
From the graph lnb versus 1/Tp and ln (b  T2
p ) versus 1/Tp,
the signals m/z scale up to 200 amu, time for each cycle 100 s.
taking into account the slope and the intercept, the pre-
exponential factor and activation energy of the crosslinking
3. Results and discussion
reactions are obtained.
The thermal stability of the crosslinked epoxy resins were
thermo gravimetrically analyzed using a STA 449 F1 Jupiter 3.1. Curing of epoxy resin with HA as hardeners
apparatus (Netzsch-Germany), coupled to a Vertex 70 spec-
trophotometer for FT-IR analysis and Aëolos QMS 403C The possible crosslinking reaction between DGEBA epoxy
mass spectrometer (Netzsch-Germany) for the mass spectro- resin and hydroxy acids (TEBAC as catalyst) is presented in
scopic analysis of the evolved gases. Samples up to 15 mg, Fig. 1. In the presence of TEBAC, the acid is deprotonated
placed in Al2O3 crucibles were thermal degraded at the heating and occurs an anioncarboxylate as a nucleophilic agent. Next,
rates of 5, 7.5, and 10 °C min1 under nitrogen atmosphere in the esterification reaction takes place by the attack of the
the temperature range between 25 and 600 °C. Initially, the anioncarboxylate to the epoxide ring, preventing thus the
kinetic parameters with Friedman (FR) and OFW methods homopolymerization of epoxy resin. (Matejka and Dusek,
included in the ‘‘Thermokinetics-3” software (Netzsch ‘‘Ther 1986). In our case, TEBAC was used as catalyst, but is it
mokinetics-3”, version 2008.05), were calculated. Based on embedded in the final product. This type of reaction is con-
the activation energy variation of degradation versus conver- firmed by FT-IR analysis. As it can be seen in Fig. 2 (FT-IR
sion degree (obtained with FR method) the mode in which spectrum for DGEBA/TA and DGEBA/CA) the increasing
thermal degradation occurs can be determined. Choosing the of the value of the absorption band located at 3411 cm1 con-
most probable kinetic model for the thermal degradation, firms the formation of the new hydroxyl groups due to the
the obtaining of the kinetic parameters can be made using reaction between epoxy and carbonyl groups. Also, this is con-
the ‘‘Multivariate regression non-linear” method from the firmed by the disappearance of the signal from 915 cm1 speci-
above software. The identification of the chemical structures fic to epoxy ring and the appearance of the new signal located
of the gases evolved in the degradation time was made with at 1741 cm1 specific to C‚O group from the new ester
FT-IR analysis and mass spectroscopic analysis. The resulted groups, as well as the signal located at 1181 cm1 (CAOAC
gases by thermal decomposition are transferred through a line group) specific to ester group which overlaps by the CAOAC

Fig. 1 The possible mechanism of DGEBA the crosslinking with CA.

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
4 N. Tudorachi, F. Mustata

the heating rate the peak temperature is shifted to higher val-


ues, this fact can be attributed both to the kinetic considera-
tions and to the delay of the instrument response (Rosu
et al., 2001; Roudsari et al., 2014; Tripathi et al., 2015) observ-
able degradation using Eqs. (2) and (3), the kinetic parameters
of curing reaction can be calculated. The different values of the
activation energy obtained for the same sample is due to the
different approximation methods used to solve the two equa-
tions. For DGEBA/CA sample, these values ranges from 61
and 67 kJ mol1 and for DGEBA/TA sample varies between
65 and 71 kJ mol1. This fact is in concordance with the liter-
ature for DGEBA crosslinked with acids as hardeners
(Mustata and Tudorachi, 2016). For non-catalyzed samples,
the shift of the exothermic peak (Fig. 4) to the higher temper-
atures justifies the use of the catalyst in the crosslinking pro-
cesses. The variation of activation energies versus conversion
degree calculated with FR method (Netzsch ‘‘Thermokinet
ics-3”, version 2008.05) (Fig. 5) indicates how the crosslinking
Fig. 2 FT-IR spectra of: (a) DGEBA, (b) (DGEBA/CA), and (c) process occurs. The fact that the shape of activation energies
(DGEBA/TA). curves presents maxima and minima suggests that the
crosslinking processes are complex processes (Worzakowska,
group of ether type from epoxy resin. Because of the very sim- 2007; Edelmann et al., 2007). The most probable kinetic model
ilar chemical structure of the two acids, the IR spectra of the and the kinetic parameters can be obtained using the ‘‘Multi-
crosslinked products are almost identical. The only observable variate regression non-linear” method from Netzsch
difference is located in the 850–960 cm1 range, the peaks for ‘‘Thermokinetics-3” software. Based on the initial values of
DGEBA/TA crosslinked product in this region being larger Ea and pre-exponential factor obtained with Friedman
than that of DGEBA/CA crosslinked product. These peaks method, the multivariate linear regression software resolve of
can be attributed to ether linkages, which may be more the numerical differential equations, using 16 different kinetic
because TA possesses two hydroxyl groups. In Fig. 3 are pre- models included in program. It may determine the most prob-
sented the DSC curves recorded at three heating rates for able global kinetic model based on differences between the cal-
DGEBA/HA samples (in presence of TEBAC) which show a culated and experimental data. The software uses a modified
single exothermic peak for each heating rate. This suggests that Marquardt procedure using Runge–Kutta analysis. The calcu-
reactions between carboxylic protons and epoxy ring take lations conducted with multivariate linear regression software
place simultaneously at all two or three carboxyl groups. With were based on the experimental data collected at three heating
the temperature increase the crosslinked products are rates and were performed by software on the conversion range
obtained. The peak temperature values for both systems are of 0.15 and 0.85. The next curing reaction mechanisms with
presented in Table 1. They grow with increasing of heating successive reactions were chosen (see Table 2):
rates and are different for each system. With the increase of For DGEBA/TA sample:
A-1 ! B-2 ! C-3 ! D with t : f; f; Bna; An; Fn reaction code

where: A are the initial reactants, B and C are the intermediate


products and D is the final crosslinked product, t:f,f; represent
the three-steps successive reaction schemes and 1, 2, 3 denote
the reaction steps.
And for DGEBA/CA sample:
A-1 ! B-2 ! C with d : f; CnB; An; reaction code
where the meanings are the same as above.
The conversion functions are:

– expanded Prout–Tompkins equation

Bna : fðaÞ ¼ ð1  aÞn aa ð4Þ

where: n - is the reaction order, a - is the degree of autocat-


alytic reaction and a - is the conversion degree)
– reaction order nth with autocatalysis model CnB
Fig. 3 Simulated and experimental DSC curves of (a) DGEBA/
Cn : fðaÞ ¼ ð1  aÞn ð1 þ Kcat  aÞ;
CA and (b) DGEBA/TA at: (j) 5 °C min1, (d) 10 °C min1 and
(N) 15 °C min1 (symbols represent the simulated curves and lines where: Kcat is the autocatalytic constant, n is the reaction order
represent experimental curves). and a is the conversion degree.

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
Curing and thermal degradation of diglycidyl ether 5

Table 1 Kinetic parameters obtained from DSC scan (stoichiometric epoxy/carboxylic proton ratio = 1/0.75).
Resin system (w/w) Heating rate (°C min1) Activation energy of curingb (kJ mol1) Frequency factorc lnA (min1)
5 10 15 Method
TMa TMa TMa EOzawa EKissinger
DGEBA/CA 112 126 133 67.32 60.79 10.78
DGEBA/TA 120 135 140 71.21 64.49 11.45
a
TM - maximum peak temperature, °C.
b
EOzawa, EKissinger - activation energies of curing reaction calculated with Eqs. (1) and (2).
c
Calculated with Kissinger equation.

Table 2 Kinetic and statistics parameters determined after


non-linear regression for the most probable mechanism of
crosslinked process by applying a kinetic model in two and
three-steps, with consecutive reactions on the temperature
interval 50–180 °C.
Parametersa DGEBA/TA DGEBA/CA
t:f,f; Bna,An,Fn d:f; CnB,An
mechanism scheme mechanism scheme
A-1 ! B-2 ! C-3 ! D A-1 ! B-2 ! C
E1/kJ mol1 90 76
log A1/s1 8.12 8.28
n1 1.11 1.26
a 0.12 3.04
Fig. 4 Experimental DSC curves of (a) DGEBA/CA and (b)
E2/kJ mol1 59 56
DGEBA/TA obtained at 10 °C min1 in absence of catalyst. log A2/s1 10.77 10.51
n2 0.44 0.824
E3/kJ mol1 101 –
log A3/s1 9.96 –
n3 0.19 –
follReact 1 0.402 0.184
follReact 2 0.313 –
Fexp 1.00 1.00
Fcrit-0.95. 1.27 1.23
t-critical (0.95;457) 1.9637 1.961
correl-coeff 0.9580 0.9955
E1, E2, E3 – activation energy of degradation for each step; log (A1,
A2, A3) – pre-exponential factor for each step; dimension 1 – is the
diffusion coefficient; n1, n2, n3 – reaction order; follReact 1 share of
reaction step 1 (A ? B), follReact 2 share of reaction step 2 (B ?
C), and follReact 3 share of reaction step 3 (C ? D) in the total
P
mass loss, is given by 1  ðfollReactÞ.

Fig. 5 Dependence of crosslinking activation energy and pre- with consecutive reactions of t:f,f; Bna, An, Fn (sample
exponential factor versus the conversion degree for: (a) DGEBA/ DGEBA/TA) and d:f; CnB, An, (Sample DGEBA/CA) types,
CA and (b) DGEBA/TA (calculated with Friedman method). are in good concordance with the experimental data, suggest-
ing that the kinetic models fairly describe them. Thus, for
– nth reaction order, model, the DGEBA/TA sample, the activation energies varies between
59 and 101 kJ mol1 with correlation coefficient of 0.9580 and
Fn : fðaÞ ¼ ð1  aÞn ; ð5Þ for DGEBA/CA sample between 56 and 76 kJ mol1 with cor-
relation coefficient of 0.9955.
where n is the reaction order.
– Avrami–Erofeev reaction model
3.1.1. Degree crosslinking of epoxy resins
An : fðaÞ ¼ nð1  aÞ½ lnð1  aÞðn1Þ=n ð6Þ
The gel content of the crosslinked of epoxy resins was obtained
where n is a constant parameter. by extraction for one day in xylene/DMF mixture. This has
Using the data presented in Table 3, the recalculated curves value of 92 for DGEBA/TA sample and 95 for DGEBA/CA
(Fig. 3) for the reaction models taking place into three-steps sample, as a consequence of the different number of reactive

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
6 N. Tudorachi, F. Mustata

Table 3 The thermal parameters obtained by TG analysis at three heating rates for DGEBA/HA systems.
Sample Heating rate Degradation stage Tonset Tpeak W T10 T20 T50 GS
°C min1 °C °C % °C °C °C °C
DGEBA/TA 5 I 184 – 4.10 299 334 375 –
II 264 – 9.75
III 332 369 76.12
Residue 10.03
7.5 I 193 – 3.42 315 346 384 –
II 281 – 9.95
III 344 380 75.45
Residue 11.18
10 I 195 – 2.78 320 356 393 317
II 287 324 10.20 429
III 350 392 75.72
Residue 11.30
DGEBA/CA 5 I 317 397 89.21 323 349 386 –
Residue 10.79
7.5 I 345 391 88.16 335 361 396 –
residue 11.84
10 I 346 398 88.86 341 365 399 421
Residue 11.14
Tonset – the temperature at which the thermal degradation start.
Tpeak – the temperature at which the degradation rate is maximum.
T10, T20, T50 – the temperatures corresponding to 10 wt%, 20 wt%, 50 wt% mass losses.
TGS – temperature at which the maximum amount of gases, was released (Gram-Schmidt curve).
W – residual mass at 600 °C.

groups in the chemical structure of the two hydroxy acids.


Thus, fewer monomers will be extracted from the DGEBA/
CA sample, because the penetration of solvents into a more
crosslinked sample will be more difficult.

3.2. TG/FT-IR/MS characterization of the cured products

The thermal behavior of the cured products was estimated


using simultaneous TG/FT-IR/MS analysis. In Fig. 1 is pre-
sented the most probable chemical structure of the crosslinked
DGEBA/CA system. In Fig. 6, are shown TG and DTG curves
for these systems registered at three heating rates (5, 7.5,
10 °C min1) under nitrogen atmosphere. From these
curves, the main parameters of the degradation process:
Tonset - temperature for initial decomposition, Tpeak - temper-
ature at maximum decomposition, T10 - temperature at
10 wt% weight loss, T20 - temperature at 20 wt% weight loss, Fig. 6 TG and DTG curves recorded at 5, 7.5 and 10 °C min1
T50 - temperature at 50 wt% weight loss, TGS - the maximum for: (a) DGEBA/CA and (b) DGEBA/TA, (symbols represent the
temperature of Gram-Schmidt curve from IR spectrum and W experimental values and lines represent the calculated values).
- weight loss at 600 °C) are summarized in Table 3. As it can be
seen from Fig. 6, with the increase of the heating rate TG
curves are shifted to higher temperatures. This phenomenon DGEBA/TA sample. For the same heating rate the peak tem-
is often found in this kind of determinations, because the sam- perature has up to 30 °C higher values. As can be seen,
ple temperature is proportionally exceeded by the furnace tem- DGEBA/CA presents only one thermal degradation process
perature due to their thermal inertia at high heating rate (Rosu with significant mass losses (about 90%), while DGEBA/TA
et al., 2011; Mustata and Tudorachi, 2010). As it can be seen in presents three thermal degradation processes with mass loss
Fig. 6 and in Table 3, the thermal decomposition begin near for each degradation step: (I = 2.78–4.10%, II = 9.75–
200 °C for DGEBA/TA system and the weight loss does not 10.20%, III = 75.45–76.12%). This confirms that DGEBA
exceed 5 wt%. For DGEBA/CA system the thermal decompo- resin was almost completely crosslinked with CA, whereas in
sition begins near 300 °C. Also, it can be seen that DGEBA/ the case of TA the crosslinking degree is much lower. This
CA sample shows only observable degradation peak and the may be a consequence of the higher number of carboxyl
degradation onset temperature has much higher values than groups from CA. For both samples the residual mass at

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
Curing and thermal degradation of diglycidyl ether 7

600 °C was about 10–11%. The difference between measured – reaction order nth model, with Fn:
temperatures for 10 and 50 wt% mass losses for both samples, fðaÞ ¼ ð1  aÞn ð10Þ
is preserved. The experimental data were processed using
‘‘Thermokinetics 3” software. Initially, this software allows where n is the reaction order and a is the conversion degree;
the calculation of the kinetic parameters using FR and OFW The calculations were effectuated for a greater number of
model free methods. These methods permit to calculate the mechanisms from which were selected the following successive
variation of the activation energy of the thermal degradation processes which had place in three or four steps:
versus the conversion degree. The data from Table 3, the shape
A-1 ! B-2 ! C-3 ! D
of DTG curves (Fig. 6) and variation of activation energies
(Fig. 7) indicate that the thermal degradation processes are and
complex processes (Worzakowska, 2007; Edelmann et al., A-1 ! B-2 ! C-3 ! D-4 ! E
2007). The shape of DTG and Ea curves suggests the entire
process of degradation is governed by at least two different with the following types of the processes schemes: t:f,f; D3,Fn,
reaction mechanisms. These processes can take place in more An (the intermediate solid products are A, B, C and D is the
steps and can run as consecutive, parallel or competitive pro- solid residue) and q:f,f,f; D4,An,An,Fn, (the intermediate solid
cesses. The identification of the thermal degradation mecha- products are A, B, C, D and E is solid residue), the numbers 1,
nisms and the calculation of the corresponding kinetic 2, 3, 4 representing the reaction steps.
parameters, were made with isoconversional differential Their choice was made using the greatest statistic coeffi-
method of Friedman and ‘‘Multivariate regression non- cients [the experimental F-value (Fexp), Fcrit (0.95) and corre-
linear” method from ‘‘Thermokinetics-3” software based on lation coefficients]. In Table 4 the obtained kinetic and statistic
the data recorded at different heating rates (Netzsch ‘‘Thermo parameters for the above models, are listed. Based on the above
kinetics-3”, version 2008.0519; Opfermann, 2000). The soft- mechanisms and kinetic parameters listed in Table 4, the recal-
ware has done the calculations based on the experimental data culated TG curves on the interval 200–600 °C show a good
collected at three heating rates on the conversion range of 0.15 agreement with experimental data, suggesting that the theoret-
and 0.85. ical models fit very well with the real phenomenon (Fig. 6). As it
The next conversion Eqs. (3)–(6) for one single step was can be seen in Table 4 for some degradation stages, the reaction
used: order greater than 1 was recorded, suggesting that during the
degradation process may appear some degradation products
– three-dimensions diffusion (Jander’s type) D3: with high molecular weight (Rosu et al., 2011). The value of
reaction orders less than 1 suggests that at this thermal degra-
fðaÞ ¼ 1:5ð1  aÞ0:666 =½1  ð1  aÞ0:333  ð7Þ dation stage, the majority of volatile products are low molecu-
lar weight products. The obtained activation energies and pre-
– three-dimensions diffusion (Ginstling-Brounstein type) D4: exponential factors (Table 4) are different in function of the
chemical structure of curing agents. DGEBA/TA activation
1
fðaÞ ¼ 1:5=½ð1  aÞ0:333  1 ð8Þ energy values between 61 and 234 kJ mol1 while for
DGEBA/CA between 55 and 314 kJ mol1. This fact is in con-
– Avrami–Erofeev reaction model An: cordance with the chemical structure of CA which has three
fðaÞ ¼ nð1  aÞ½ lnð1  aÞðn1Þ=n ð9Þ carboxyl groups and can induce higher degree of crosslinking
with increase of thermal stability, consequently. Also, the con-
where n is a constant parameter. clusion that CA vs. TA can induce a higher crosslinking degree
was confirmed by the Shore and Vickers hardness tests, as well
as the gel percentage. Also, the conclusion that CA vs. TA can
induce a higher crosslinking degree was confirmed by Shore
and Vickers hardness tests as well as the percentage of gel. If
considering Tpeak,W600 parameters and activation energies of
the thermal decomposition as thermal stability criteria, it can
be seen that only DGEBA/CA sample has the same degree of
thermal stability in comparison with DGEBA crosslinked with
p-aminobenzoic acid or with dicarboxylic aromatic acids [Rosu
et al., 2011; Mustata et al., 2015; Mustata and Tudorachi,
2016]. By comparison (Fig. 8) it can be seen that for the samples
obtained in the absence of the catalyst, several degradation
peaks occur with the maxima at lower temperatures than those
obtained in the presence of the catalyst. This can be a conse-
quence of a lower crosslinking degree in absence of the catalyst.

3.3. Chemical composition of the evolved gases

Fig. 7 Dependence of thermal decomposition activation energy In Fig. 9 is shown the most likely mechanism of thermal
and pre-exponential factor versus conversion degree for: (a) destruction of the cured sample. The chemical structure of
DGEBA/TA and (b) DGEBA/CA (calculated with Friedman gases that result from thermal degradation was identified using
method). FT-IR and MS analysis. In Fig. 10 are shown the two and

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
8 N. Tudorachi, F. Mustata

Table 4 Kinetic and statistics parameters determined after non-linear regression for the most probable mechanism of thermal
degradation process of the cured products by applying a kinetic model in three and four-steps, with consecutive reactions on the
temperature interval 50–600 °C.
Parametersa DGEBA/TA DGEBA/CA
t:f,f; D3,An,Fn q:f,f,f; D4,An,An,Fn
mechanism scheme mechanism scheme
A-1 ! B-2 ! C-3 ! D A-1 ! B-2 ! C-3 ! D-4 ! E
E1/kJ mol1 61 81
log A1/s1 2.54 3.31
E2/kJ mol1 133 55
log A2/s1 8.19 30.97
n2 0.916 2.97
E3/kJ mol1 234 168
log A3/s1 15.10 12.03
n3 2.99 0.232
E4/kJ mol1 – 314
log A4/s1 – 23.08
n4 – 2.734
follReact 1 0.1205 0.1010
follReact 2 0.7408 0.1167
follReact 3 – 0.2974
Fexp 1.00 1.00
Fcrit-0.95. 1.14 1.14
t-critical (0.95;457) 1.955 1.955
correl-coeff 0.999627 0.999602
E1, E2, E3, E4 – activation energy of degradation for each step; log (A1, A2, A3, A4) – pre-exponential factor for each step; dimension 1 - is the
diffusion coefficient; n2, n3, n4 – reaction order; follReact 1 share of reaction step 1 (A ? B), follReact 2 shareP of reaction step 2 (B ? C),
follReact 3 share of reaction step 3 (C ? D) and share of step 4 (D ? E) in the total mass loss, is given by 1  ðfollReactÞ:

of 424 °C, the intensity of the absorption bands shows a max-


imum. In Fig. 10b can be observed a relative large number of
signals that are located at 3655, 3254, 2963, 2439, 2354, 1729,
1605, 1510, 1250, 1180, 1039 and 827 cm1. The significance of
these signals was assigned based on visual observation and on
the data from literature (Silverstein et al., 2005). Thus, the sig-
nals located between 3200 and 3700 cm1 can be assigned to
water vapors and alcohols which can appear at the thermal
degradation of the secondary hydroxyl, ester or ether groups.
The signals present at 3040, 2800 cm1 and 1400, 1300 cm1
can be assigned to the vibration of CH, CH2 and CH3 groups
located in the chemical structure of saturated and unsaturated
aliphatic hydrocarbons or aromatic compounds, which origi-
nate from the degradation of acids or DGEBA moieties. The
strongest signal present at 2354 cm1 and can be certainly
attributed to carbon dioxide, which can occur as result of ester
Fig. 8 TG and DTG experimental curves recorded at
group degradation (Worzakowska and Scigalski, 2014). Also,
10 °C min1 for: (N) DGEBA/CA and () DGEBA/TA, cross-
at 2180 cm1 is present one small signal which suggests that
linked without catalyst.
an amount of carbon monoxide can appear. It would be
derived from the degradation of the hydroxyl groups located
three-dimensional FT-IR spectra (2D and 3D spectra) for the on the hydroxy acid molecule. The signal located at
gases resulted at the thermal degradation of DGEBA/TA sam- 1729 cm1 can be assigned to the C‚O groups from acids,
ple registered at 10 °C min1 heating rate the temperature esters, anhydrides, aldehydes products occurring at the degra-
range of 40–600 °C. 2D and 3D spectra of the evolved gases dation of ester groups formed after cross-linking reaction. The
at the degradation of DGEBA/CA sample, recorded in the signals located at 1605, 1510 and 829 cm1 are specific for the
same conditions as above are similar and are not presented aromatic derivatives such as phenol, benzene, toluene, a-
from the reason of brevity. In FT-IR 3D spectrum is shown methyl styrene etc. resulted at the cracking of DGEBA moi-
the absorbance versus the wavelength and temperature, while eties. Also, for these substances the signal located at
in 2D spectrum is plotted the absorbance versus the wave- 3040 cm1 which is specific to the C‚C aromatic bonds, could
length at the maximum temperature from Gram Schmidt plot be assigned. The signals situated at 1250 and 1180 cm1 can be
(424 °C). As it can be seen in 3D spectrum at the temperature assigned to CAO group, located in ester type gases.

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
Curing and thermal degradation of diglycidyl ether 9

Fig. 9 The possible mechanism of cured DGEBA/TA thermal degradation.

signals are located up to m/z = 70 and are specific to more spe-


cies. The identification of the chemical structure of gases
resulted after thermal degradation was made using data from
NIST MS library on the base of molecular weight of ionic frag-
ments (http://webbook.nist.gov/chemistry/name-ser.html).
The ionic fragments located up to 50 amu can be assigned to
water (m/z = 18, 17), carbon dioxide (m/z = 44, 28, 16, 12),
saturated and unsaturated aliphatic hydrocarbons [methane
(m/z = 16, 15, 14, 13, 12), propane (m/z = 44, 43, 39, 29),
ethylene (m/z = 28, 27, 26, 24, 14) propene (m/z = 42, 41,
40, 39, 38, 37, 27, 26, 15), acetylene (m/z = 26, 25, 24, 13)].
These ionic fragments may be derived by degradation of ali-
phatic moieties of the two-components. The fragments which
appear in the range of 50–60 amu can be allocated to aldehyde,
ketone, ester or acid type products, and may originate espe-
cially from degradation of the hardener moieties [2-propenal
(m/z = 56, 55, 37, 29, 28, 27, 26, 25), acetone (58, 43, 42,
27,15), formic acid (m/z = 46, 45, 42, 29, 28, 17), acetylalde-
hyde (44, 43, 42, 29, 15), acetic acid (m/z = 60, 45, 43, 42,
29, 15), methyl formate (m/z = 60, 32, 31, 29, 15)]. The above
substances can result mainly by degradation of the ester link-
age obtained from the reaction between epoxide groups and
carboxyl groups of the hydroxy acids. The signals with values
over 70 amu can be assigned to the components of aromatic
type [benzene (m/z = 78, 77, 50, 39), phenol (m/z = 94, 66,
65, 39), toluene (92, 91, 65, 51, 39) isopropylbenzene (m/
z = 120, 105, 103, 91, 79, 77, 51, 39)] obtained from the ther-
mal degradation of DGEBA moieties. Such gaseous products
have been identified at the degradation of DGEBA thermoset-
ting resins (Mustata et al., 2011). In Fig. 11, are shown the
evolved gases profiles on the temperature range 100–600 °C
for some MS signals (for clarity were chosen for presentation
only few signals which appear at the thermal degradation of
DGEBA/TA sample) specific to water, carbon dioxide,
methane, propane, acetylene, acetic acid, benzene, phenol, iso-
propyl benzene. For the DGEBA/CA sample, the evolved
gases profile is relatively similar. According to the degradation
model from Fig. 9 correlated with Figs. 10 and 11, it can be
considered that at lower temperatures (about 200 °C) takes
place the breakdown of the secondary hydroxyl and ester
groups, with the obtaining of water, carbon dioxide and car-
boxyl products (acids, esters, ketones, aldehydes), while at
higher temperatures (about 400 °C) aromatic products (ben-
zene, phenol, toluene) are resulted.
Fig. 10 Stacked plot diagram (a), FT-IR spectrum (b) and MS
spectrum (c) of evolved gases obtained at 424 °C for DGEBA/TA. 3.4. Morphological study of the cured samples

The assignments in the IR spectra are supported by MS In Table 5 is shown the Shore and Vickers hardness for the
representations from Fig. 10c. From Fig. 10c (for brevity only samples crosslinked in the presence and absence of TEBAC.
MS spectrum for DGEBA/TA is presented) the majority of As can be seen, the DGEBA/CA samples have higher hardness

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
10 N. Tudorachi, F. Mustata

Fig. 11 The variation of ion current intensity with temperature at heating rate of 10 °C min1 for the main signals of the evolved gases at
DGEBA/TA sample.thermal degradation.

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
Curing and thermal degradation of diglycidyl ether 11

Table 5 Hardness of the epoxy crosslinked resin.


Sample DGEBA/CA with DGEBA/TA with DGEBA/CA without DGEBA/TA without
TEBAC TEBAC TEBAC TEBAC
Shore hardness 93 82 55 22
Vickers hardness (VH 15.53 14.21 10.21 –
units)

compared to DGEBA/TA samples crosslinked in presence of higher degree of crosslinking of DGEBA/CA sample, so a
the catalyst. The same trend is maintained in the case of cross- more rigid structure resulted with, a brittle behavior. In
linked samples in absence of the catalyst. The toughness of Fig.12b is shown the surface fracture for DGEBA/TA sample
DGEBA/HA systems can be correlated with the chemical com- where it can be observed that the surfaces present the fracture
position of the hardener. The aspect of the fracture surfaces lines with more ridges and tortuous cracks. These forms of the
(SEM micrographs-magnification of 5000) obtained after cool- fracture lines show that, due to chemical structure of DGEBA/
ing at liquid nitrogen temperature for DGEBA/HA cross- TA sample (lower degree of crosslinking), the breaking has
linked samples are shown in Fig. 12. The appearance of the ductile characteristics. The gel content also confirms the differ-
fractured surface allows making a correlation between the ent crosslinking degrees between DGEBA/CA and DGEBA/
morphology and toughness of samples. In Fig.12a for TA samples. Similar characteristics were observed on the frac-
DGEBA/CA sample, the cracks of the surface appear as ture surface of DGEBA thermoset resin crosslinked with male-
smooth glass in which the fracture lines are parallel lines with- opimaric acid (Mustata et al., 2014).
out any proof of deformation, indicating that the thermoset
product has a poor impact strength (Mustata et al., 2013b). 4. Conclusion
This is due to the chemical structure of CA which induces a
Curing and thermal degradation mechanisms of diglycidyl
ether of bisphenol A crosslinked with natural hydroxy acids
were studied using DSC and simultaneous TG/FT-IR/MS
techniques. The kinetic parameters of the crosslinking reac-
tions depends on the chemical structure of the curing agents
and have values for activation energies which ranges between
61 and 67 kJ mol1 for DGEBA/CA sample and between 65
and 71 kJ mol1 for DGEBA/TA sample.
The Ea values of thermal degradation calculated with the
Friedman isoconversional method vary with the conversion
degree. On the other hand, the form of DTG curves and ther-
mal parameters suggest that the apparent mechanism of ther-
mal decomposition of the cured samples is a complex
phenomenon and occurs in three or four stages depending
on the chemical structure of the curing agent. The kinetic
parameters suggest that DGEBA/CA sample is more thermally
stable than DGEBA/TA sample. This can be justified on the
basis of chemical structure of CA with three carboxyl groups
which lead to a strong crosslinked product. Gaseous products
which appear as a result of the thermal destruction, identified
FT-IR and MS analysis, are mainly water, carbon dioxide,
hydrocarbons and acids derivatives. The fractured surfaces
of DGEBA/TA sample present the fracture lines with more
ridges and tortuous cracks, while the crack surface of
DGEBA/CA sample presents a smooth glassy surface with
the fracture parallel lines without any proof of deformation.

References

Alam, M., Akram, D., Sharmin, E., Zafar, F., Ahmad, S.S., 2014.
Vegetable oil based eco-friendly coating materials: a review article.
Arab. J. Chem. 7 (4), 469–479.
ASTM D 1652-04, Standard test method for epoxy content of epoxy
resins.
Das, G., Karak, N., 2009. Epoxidized Mesua ferrea L. seed oil-based
Fig. 12 SEM micrographs of: (a) DGEBA/CA; (b) DGEBA/TA reactive diluent for BPA epoxy resin and their green nanocompos-
at magnification of 5000. ites. Prog. Org. Coat. 66, 59–64.

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008
12 N. Tudorachi, F. Mustata

Edelmann, M., Gedan-Smolka, M., Heinrich, G., Lehmann, D., 2007. resins. Curing kinetics, thermal properties and morphology. Com-
Thermokinetic analysis of two-step curing reactions in melt Part I. posites: Part B 55, 470–478.
Investigation of low molecular model systems. Thermochim. Acta Mustata, F., Tudorachi, N., Bicu, I., 2014. The curing reaction of
452, 59–64. epoxidized methyl esters of corn oil with Diels-Alder adducts of
Ellis, B., 1993. In: Chemistry and Technology of Epoxy Resins, vol. 1. resin acids. The kinetic study and thermal characterization of
Blakie Academic Professional, Glasgow. crosslinked products. J. Anal. Appl. Pyrol. 108, 254–264.
Flynn, H.J., Wall, L.A., 1966. General treatment of the thermo- Mustata, F., Tudorachi, N., Bicu, I., 2015. The kinetic study and
gravimetry of polymers, J. Res. Nat. Bur. Stand. A Phys. Chem. 70, thermal characterization of epoxy resins crosslinked with amino
487–523. carboxylic acids. J. Anal. Appl. Pyrol. 112, 180–191.
Galego, N., Vazquez, A., Riccardi, C.C., Williams, R.J.J., 1992. Citric Mustata, F., Tudorachi, N., 2016. Thermal behavior of epoxy resin
acid–diethylenetriamine salts as latent curing agents for epoxy cured with aromatic dicarboxylic acids. J. Therm. Anal. Calorim.
resins. J. Appl. Polym. Sci. 45 (4), 607–610. 125, 97–110.
Galego, N., González, F., Vazquez, A., 1996. Curing of epoxy resins Netzsch ‘‘Thermokinetics-3”, version 2008.05.
with citric acid–piperazine salt. Polym. Int. 40 (3), 213–218. Opfermann, J., 2000. Kinetic analysis using multivariate non-linear
http://webbook.nist.gov/chemistry/name-ser.html. regression. I. Basic concepts. J. Therm. Anal. Calorim. 60, 641–658.
Jaillet, F., Desroches, M., Auvergne, R., Boutevin, B., Caillol, S., Ozawa, T., 1976. A modified method for kinetic analysis of thermo-
2013. New biobased carboxylic acid hardeners for epoxy resins. analytical data. J. Therm. Anal. 9, 369–373.
Eur. J. Lipid Sci. Technol. 115, 698–708. Patil, D.M., Phalak, G.A., Mhaske, S.T., 2017. Synthesis of bio-based
Kissinger, H.F., 1957. Reaction kinetics in different thermal analysis. epoxy resin from gallic acid with various epoxy equivalent weights
Anal. Chem. 29, 1702–1706. and its effects on coating properties. J. Coat. Technol. Res. 14 (2),
Lee, H., Neville, K., 1972. Handbook of Epoxy Resins. McGraw-Hill 355–365.
Inc., New York. Park, S.-J., Jin, F.-L., Lee, J.-R., 2004. Thermal and mechanical
Li, F., Xiao, F., Moon, K.-S., Wong, C.P., 2006. Novel curing agent properties of tetrafunctional epoxy resin toughened with epoxidized
for lead-free electronics: amino acid. J. Polym. Sci.: Part A: Polym. soybean oil. Mater. Sci. Eng., A 374, 109–114.
Chem. 44, 1020–1027. Rosu, D., Mustata, F., Cascaval, C.N., 2001. Investigation of the
Liu, X.Q., Huang, W., Jiang, Y.H., Zhu, J., Zhang, C.Z., 2012. curing reactions of some multifunctional epoxy resins using
Preparation of a bio-based epoxy with comparable properties to differential scanning calorimetry. Thermochim. Acta 370, 105–110.
those of petroleum-based counterparts. Express Polym. Lett. 6, Rosu, D., Rosu, L., Brebu, M., 2011. Thermal stability of silver
293–298. sulfathiazole-epoxy resin network. J. Anal. Appl. Pyrol. 92, 10–18.
Matejka, L., Dusek, K., 1986. Specific features of the kinetics of Roudsari, G.M., Mohanty, A.K., Misra, M., 2014. Study of the curing
addition esterification of epoxide with the carboxyl group. Polym. kinetics of epoxy resins with biobased hardener and epoxidized
Bull. 15, 215–221. soybean oil. ACS Sustain. Chem. Eng. 2, 2111–2116.
May, C.A., 1988. Epoxy Resins-Chemistry and Technology. Marcel Silverstein, R.M., Webster, F.X., Kiemle, D.J., 2005. Spectrometric
Dekker Inc., New York. Identification of Organic Compounds. Wiley & Sons Inc.,
Miyagawa, H., Mohanty, A.K., Misra, M., Drzal, L.T., 2004. Hoboken.
Thermophysical and impact properties of epoxy containing epox- Tripathi, M., Kumar, D., Rajagopal, C., Kumar Roy, P., 2015. Curing
idized linseed oil. 1. Macromol. Mater. Eng. 289, 629–635. kinetics of self-healing epoxy thermosets. J. Therm. Anal. Calorim.
Mustata, F., Tudorachi, N., 2010. Epoxy resins crosslinked with rosin 119, 547–555.
adduct derivatives. Crosslinking and thermal behaviors. Ind. Eng. Vyazovkin, S., Chrissafis, K., Di Lorenzo, M.L., Koga, N., Pijolat,
Chem. Res. 49, 12414–12422. M., Roduit, B., Sbirrazzuoli, N., Suñol, J.J., 2014. ICTAC Kinetics
Mustata, F., Tudorachi, N., Rosu, D., 2011. Curing and thermal Committee recommendations for collecting experimental thermal
behavior of resin matrix for composites based on epoxidized analysis data for kinetic computations. Thermochim. Acta 590, 1–
soybean oil/diglycidyl ether of bisphenol A. Compos. B 42, 1803– 23.
1812. Worzakowska, M., 2007. The kinetic study of the curing reaction of
Mustata, F., Tudorachi, N., Bicu, I., 2013a. Bio-based epoxy matrix mono- and di-epoxides obtained during the reaction of divinyl-
from diglycidyl ether of bisphenol A and epoxidized corn oil, benzene and hydrogen peroxide with acid anhydrides. Polymer 48,
crosslinked with Diels-Alder adduct of levopimaric acid with 1148–1154.
acrylic acid. Ind. Eng. Chem. Res. 52, 17099–17110. Worzakowska, M., Scigalski, P., 2014. Synthesis and thermal behavior
Mustata, F., Tudorachi, N., Bicu, I., 2013b. Thermosetting resins of linear neryl diesters in inert and oxidative atmosphere. J. Therm.
obtained via sequential photo and thermal crosslinking of epoxy Anal. Calorim. 115, 783–793.

Please cite this article in press as: Tudorachi, N., Mustata, F. Curing and thermal degradation of diglycidyl ether of bisphenol A epoxy resin crosslinked with natural
hydroxy acids as environmentally friendly hardeners. Arabian Journal of Chemistry (2017), http://dx.doi.org/10.1016/j.arabjc.2017.07.008

Potrebbero piacerti anche