You are on page 1of 7

Thermochimica Acta 555 (2013) 4652

Contents lists available at SciVerse ScienceDirect

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

Thermal degradation characteristics of phenolformaldehyde resins derived


from beetle infested pine barks
Yong Zhao a , Ning Yan a, , Martin W. Feng b
a
b

Faculty of Forestry, University of Toronto, 33 Willcocks Street, Toronto, ON, Canada M5S 3B3
FPInnovations, Wood Products Division, 2665 East Mall, Vancouver, BC, Canada V6T 1W5

a r t i c l e

i n f o

Article history:
Received 15 October 2012
Received in revised form
18 December 2012
Accepted 20 December 2012
Available online 3 January 2013
Keywords:
Bark
Bark extractives
Phenol-liqueed bark
Bark-derived PF resin
Thermal stability
Thermal degradation

a b s t r a c t
In this study, phenolformaldehyde (PF) resins were synthesized using both bark extractives and phenolliqueed barks from mountain pine beetle (Dendroctonus ponderosae Hopkins) infested lodgepole pine
(Pinus contorta Dougl.). The thermal stability and thermal degradation kinetics of the bark-derived PF
resins were investigated by thermogravimetric analysis (TGA). The structural changes in the bark-derived
PF resins at different stages of the thermal degradation were studied using the Fourier Transform Infrared
Spectroscopy (FTIR). Thermal stability of the post-cured bark-derived PF resins were similar to that of
the lab PF resin but differed signicantly from that of the post-cured commercial PF resin. Bark-derived
PF resin made from bark extractives differed signicantly in thermal stability and thermal degradation
kinetics from the bark-derived PF resin made from the phenol-liqueed bark. The bark-derived PF resin
made from the bark extractives showed a higher thermal stability.
2012 Elsevier B.V. All rights reserved.

1. Introduction
Phenolformaldehyde (PF) resins have been widely used as
adhesives, coatings, thermal insulation materials, and molding
compounds due to their good mechanical properties and heat resistance [1]. Generally, phenol and formaldehyde go through addition
and condensation reactions under either alkaline condition to form
cross-linked thermoset polymers or under acidic condition to form
linear thermoplastic polymers. The structure and properties of the
polymer depend highly on the reaction conditions.
Thermal stability of a thermoset polymer is highly dependent
on its structure and cross-linking density [2,3]. It is an important
property of polymer durability. Various studies have investigated
the thermal degradation mechanisms for PF resins [48]. It has
been suggested that the thermal degradation of PF resins follows
an auto-oxidization process that mainly includes three stages. In
the rst thermal degradation stage, additional cross-linking was
formed between the remaining un-reacted functional groups in the
cured PF resins. In the second thermal degradation stage, the level
of cross-linking was reduced with some degradation by-products,
such as methane, hydrogen, carbon monoxide, small oligomers,
and water. In the third thermal degradation stage, the cured resin

Corresponding author. Tel.: +1 4169468070; fax: +1 4169783834.


E-mail address: ning.yan@utoronto.ca (N. Yan).
0040-6031/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.tca.2012.12.002

network would collapse followed by the carbonization and graphitization processes [4,5].
Thermogravimetric analysis that measures a samples mass loss
as a function of temperature and/or time is a common technique for
characterizing polymer thermal degradation behavior [68]. Kinetics during the thermal degradation process has been studied to
acquire fundamental understanding of the structural changes in
phenolic resins [4]. The technique used to obtain the kinetic parameters is termed either as a differential or an integral method. Both
differential and integral method can be further classied based on
either a single heating rate or multiple heating rates method is used
[9,10]. It is believed that multiple heating rates method gives more
reliable results than the single heating rate method with a smaller
experimental error [10]. The model-free isoconversional methods are widely adopted techniques for deriving relevant kinetic
parameters using thermogravimetric (TG) and differential thermogravimetric (DTG) curves measured at different heating rates,
which make no presumption about the reaction function and reaction order [915].
Previous studies [16,17,19] have also shown that thermal
stability and thermal degradation kinetics of the PF resins are signicantly affected by the resin synthesis conditions. For example, it
was found that the activation energies of the resol resins with various molar ratios of formaldehyde to phenol decreased sharply at
rst and then remained almost constant during the thermal decomposition process. However, the activation energies of the novolac

Y. Zhao et al. / Thermochimica Acta 555 (2013) 4652

resins with some formaldehyde to phenol molar ratios were almost


unchanged in the whole degradation process, while the activation
energies of the novolac resins with other formaldehyde to phenol
molar ratios gradually decreased [19].
The inclusion of other materials, such as lignin and cellulose bers, in the PF resin synthesis has been shown to modify
thermal stability and degradation kinetics of the resulting resins
[18,20,21,24]. Ligninphenolformaldehyde (LPF) resol resins were
found to have higher thermal stability under both nitrogen and air
conditions than PF resins with different degradation characteristics
[18]. Meanwhile, the LPF novolac resins were found to be less stable at lower temperatures than PF resins, but they were comparable
to PF resins at higher temperatures [20]. The addition of cellulose
bers to PF resol resins reduced thermal resistance of the resulting
resol resins [21].
Recently, phenolic compounds from bark were used to partially
substitute petroleum-derived phenol in the PF resol resin synthesis [22,23]. In these studies, bark components were obtained by
either liquefaction of bark in phenol or extraction of bark using
a solvent. The bark-derived PF resins, with both liqueed bark
and extracted bark components, were able to have comparable
bonding strengths to a commercial PF resin. But the activation
energy and pre-exponential factors of the bark-derived PF resins
in curing differed from those of commercial PF resins, suggesting structural differences in these cured adhesives. Even though
stability of the cured adhesives is an important consideration in
certain applications, no previous study has reported the impacts of
bark components on the thermal stability and degradation kinetics
of the resulting PF resins. In this study, the thermal stability and
thermal degradation kinetics of the bark-derived PF resins were
investigated.
2. Experimental
2.1. Resin synthesis

where is the degree of conversion, f() is the reaction model, and


k(T) is represented by an Arrhenius equation shown as:

k(T ) = A exp

E
RT

(2)

where A is the pre-exponential factor; E is the activation energy; T


is the temperature; R is the ideal gas constant.
KissingerAkahiraSunose (a more accurate version of
FlynnWallOzawa) and Kissinger methods [9,11,12], which
were used previously to study thermal degradation kinetics of PF
adhesives and recommended by the Kinetics Committee of the
International Confederation for Thermal Analysis and Calorimetry
(ICTAC) for performing kinetics computation on thermal analysis
data, were used to determine the kinetic parameters of the thermal
degradation process for the bark-derived PF resins.
For the KissingerAkahiraSunose method, the expression is
given as:
ln

i
2
T,i

= const

E
RT

(3)

where i is the heating rate under ith temperature program. T,i is


the temperature at which the degree of conversion, , is reached
under ith temperature program.
For the Kissinger method, the expression is:

ln

Tp2

= ln

 AR 
E

E
RTp

(4)

where Tp is the peak temperature obtained from the DTG curves of


the tested resins at different heating rates.

2.2. Measurements of thermal stability


The LBPF, BEPF, lab PF and commercial PF resins were placed in
an oven and cured at 80 C for 48 h. The cured resin samples were
ground into ne powders of sizes that had passed through a 100mesh screen and then tested in a thermogravimetric analyzer (TGA,
Q500, TA, USA). About 10 mg of each cured sample was added to a
platinum pan and heated from room temperature to 800 C at the
heating rate of 5, 10, 15 and 20 C/min under N2 atmosphere.
2.3. Kinetic analysis
The kinetic parameters for the thermal degradation process of
the bark-derived PF resins were obtained by the thermogravimetric
analysis employing multiple heating rates. It is known that the rate
of thermal degradation can be expressed as:
d
= f ()k(T )
dt

Fig. 1. Thermal degradation curves of the phenolic resins.

Liqueed bark-phenolformaldehyde (LBPF) resin and bark


extractive-phenolformaldehyde (BEPF) resin were synthesized
according to the methods reported previously [22,23]. A laboratory
made PF resin (lab PF) without any bark components was prepared
by following the same procedures used in the synthesis of bark
derived-PF resins. All chemicals were purchased from Caledon Laboratory Chemicals, Canada, and used without further purication. A
commercial PF resin for oriented strandboard face layers was used
as a comparison.

47

(1)

2.4. FTIR
The resin samples after degradation at temperatures of 200 C,
400 C, 600 C, and 800 C were collected and analyzed using Fourier
Transform Infrared Spectroscopy (FTIR). The FTIR measurements
were carried out using a FT-IR TENSOR 27 spectrometer with
ATR attachment (Bruker Optics, USA) having a frequency range of
4000400 cm1 . Elemental analysis was carried out using a 2400
Series II Carbon Hydrogen Nitrogen (CHNS) elemental analyzer.
3. Results and discussion
3.1. Thermal stability of the bark-derived PF resins
The thermal degradation curves of the different resins are shown
in Fig. 1. When the temperature increased from room temperature
to 200 C, the weight loss of the cured resins was mainly caused

48

Y. Zhao et al. / Thermochimica Acta 555 (2013) 4652

Fig. 3. Activation energies of the tested resins at different conversion levels calculated by the KissingerAkahiraSunose method.

Fig. 2. Derivative thermal degradation (DTG) curves of the phenolic resins.

by the evaporation of moisture, dehydration, as well as the evaporation of compounds with small molecular weights. The evolution
of water could be attributed to the condensation reaction between
the residual methylol groups and phenolic OH groups. Water elimination might lead to the formation of new crosslinks [4,5].
From 200 C to 400 C, the resins underwent further degradation. The release of free phenol, aldehyde, short oligomers
and water were main contributors to the weight loss. When
the temperature increased from 400 C to 600 C, major polymer
decomposition took place. The weight loss was the results of the
degradation of the polymeric molecules and the formation of small
and volatile molecules, such as CO, CO2 , benzaldehyde and phenol, and the initial formation of char [68]. After the temperature
increased to above 600 C, the degradation continued and a charlike structure of carbon was gradually formed, generating carbon
monoxide as by product [48]. It was also considered to be the
carbonization process.
Fig. 2 clearly shows that the thermal degradation process of
the commercial PF resin differed signicantly from those of the
lab PF and bark-derived PF resins, especially when the temperature reached higher than 200 C. Two distinct derivative thermal
degradation (DTG) peaks were observed (shown in Fig. 2) in the
temperature range of 200400 C for the commercial PF resin. The
reason for this was due to the fact that a signicant amount of urea
was present in the commercial PF resin. Urea is known to degrade at
such a temperature range. Elemental analysis revealed that a nitrogen content of 13% was present in the cured commercial PF resin
and it decreased to 4% when the temperature increased from room
temperature to 400 C.
The weight losses of all the tested resins during the thermal
degradation process are shown in Table 1. The total weight loss
of the commercial PF resin from room temperature to 800 C was
higher than those of the lab PF resin and bark-derived PF resins.
After the entire thermal degradation process, the bark extractivePF resin showed the highest residual weight, followed by the lab PF
resin, liqueed bark-PF resin and commercial PF resin.
The inclusion of urea in the commercial PF was the main reason
for its higher weight loss (35.3%) than those of the lab PF (20.7%),

liqueed bark-PF resin (22.7%) and bark extractive-PF resin (18.7%)


in the temperature range of room temperature to 400 C. When the
temperature increased from 400 C to 600 C, the bark-derived PF
resins produced slightly lower weight losses in comparison with
the lab PF and the commercial PF resin. From 600 C to 800 C, the
weight loss of the commercial PF resin was the lowest among all
the test resins, at 5.0%. The lab PF and bark-extractive PF resin had
similar weight losses during this temperature range, at 8.1% and
8.7%, respectively; while the liqueed bark-PF resin had the highest
weight loss, i.e. 12.1%. It is believed that the liqueed bark components with phenolic side chains or branch structures in the liqueed
bark-PF resin were thermally stable at lower temperatures but
started to degrade at higher temperatures.
The difference of the weight loss between the bark extractive-PF
resin and liqueed bark-PF resin could be attributed to the difference in chemical composition between the liqueed bark and bark
extractives as well as the difference in the resin structures. In the
recent structural study of the bark-derived PF resins, it was found
that the liqueed bark-PF had a higher ratio of para-para/ortho-para
methylene link, a higher unsubstituted/substituted hydrogen ratio
and a higher methylol/methylene ratio than the bark extractive-PF
resin. The phenolated products in the liqueed bark had multiphenolic rings and had more reactive sites toward formaldehyde
than the bark extractives. The large molecules of the liqueed bark
and bark extractives with lower molecular mobility could make it
difcult for the resulting resins to form cross-links [27,28].

3.2. Thermal degradation kinetics


The dependence of the activation energy on the extent of
degradation provides information on how the polymer structure
affects the degradation kinetics [9,10]. The linear relationship
2 ) and
between ln(/Tp2 ) and 1/Tp (Kissinger method), ln(/T,i
1/T (KissingerAkahiraSunose method) were obtained (gures
not shown). The activation energies (E ) of the tested resins
at different stages of thermal degradation calculated by the
KissingerAkahiraSunose method from the slope is shown in
Fig. 3. Even though the pre-exponential factors were also obtained,
given the focus of the study was on the thermal stability as indicated

Table 1
Weight losses of different phenolic resins during the thermal degradation process.
Resin type

Weight loss at different temperature range (%)

Commercial PF
Liqueed bark-PF
Bark extractive-PF
Lab-PF

Total weight loss (%)

RT 200 C

200400 C

400600 C

600800 C

RT 800 C

13.2
9.2
7.5
10.3

22.1
13.5
11.2
10.4

14.9
12.8
14.0
14.8

5.0
12.1
8.7
8.1

55.2
47.6
41.4
43.6

Y. Zhao et al. / Thermochimica Acta 555 (2013) 4652

49

Fig. 4. FTIR spectra of the bark extractive-PF resin at different degradation temperatures.

Fig. 5. FTIR spectra of the liqueed bark-PF resin at different degradation temperatures.

by the activation energy parameters instead of the degradation


reaction rate, they were not included in the discussions.
In the early stage (conversion <20%), the commercial PF resin
had the lowest activation energy; it could be explained by the fact
that the commercial PF resin contained a signicant amount of urea,
which easily decomposed at lower temperatures. For all the test
resins, the activation energies generally increased before the conversion reached 60%. The increased activation energies indicated
the gradual degradation of the main networks of the resins.
The highest activation energies of the bark-derived PF resins, the
lab PF resin and commercial PF resin were observed when the conversion rate was higher than 70%. It may be due to the fact that in
the late thermal degradation stage, their structures became more
orderly and required more energy to break down. The decomposition mechanism changed when the conversion reached higher
than 70% [1820]. Degradation and formation of new cross-links
occurred at the same time [4,18,20].
The bark extractive-PF resin exhibited the highest activation
energy compared with the liqueed bark-PF resin, lab PF resin
and commercial PF resins in most of the stages of the thermal
degradation process. It indicated that the cured bark extractive-PF
resin was more thermally stable than the cured liqueed barkPF. This is also consistent with different weight losses between
the bark extractive-PF resin and liqueed bark-PF resin as shown
before. The bark components in the bark-derived PF resins did affect
their post-cured crosslink density. The differences in the chemical composition and structures between the bark extractive-PF and
liqueed bark-PF resin could contribute to their different thermal
stability and thermal degradation behaviors.
The activation energies of all the tested resins calculated by
the Kissinger method by plotting ln(/Tp2 ) and 1/Tp obtained
from their DTG curves are shown in Table 2. The activation
energy of the bark extractive-PF resin was higher than those of

Table 3
Peak assignment of the FTIR spectra of all tested resins.
Wavenumber (cm1 )

Peak assignment

33503500
28482914
2164, 2227
1740, 1650, 1600
1587
1467
1435
1353
12491253, 12031253, 1170
1050
1006
879
783
621

Phenolic OH stretch
Aliphatic CH2 asymmetric stretch
CN, NCO groups
Carbonyl groups, benzene ring
Benzene ring
Benzene ring
CH2 scissor bending
Phenolic OH in-plane deformation
Alkyl-phenol C O stretch
Aromatic CH in-plane deformation
Aromatic linked CH3 rock and/or aromatic CH
Polysubstituted aromatic ring
CH2 out-of-plane ring deformation
CH2 out-of-plane ring deformation

other resins in most of the cases. The same was observed by the
KissingerAkahiraSunose method. For the commercial PF resin,
the activation energy of the second peak was lower than the other
peaks, which could be attributed to urea. The results and trends
calculated by the Kissinger method were also consistent with the
results calculated by the KissingerAkahiraSunose method.
3.3. Structural changes of the bark-derived PF resins during
thermal degradation
The structural changes of the four types of cured phenolic resins
at different thermal degradation stages were characterized by FTIR.
The FTIR spectra of the tested resins are shown from Figs. 4 to 7.
Peaks assignments were given in Table 3 according to previous studies [5,18,26]. The intensities and bands of the functional

Table 2
Activation energies of the tested resins calculated by the Kissinger method.
Activation energy (kJ/mol)

Bark extractive-PF
Liqueed bark-PF
Commercial PF
Lab PF

Peak 1

Peak 2

Peak 3

Peak 4

Peak 5

110.96 (0.98)
102.74 (0.97)
106.71 (0.99)
97.29 (0.97)

254.75 (0.94)
206.02 (0.93)
103.21 (0.99)
263.19 (0.98)

232.96 (0.98)
222.12 (0.97)
138.84 (0.99)
219.98 (0.97)

NA
NA
118.05 (0.99)
NA

NA
NA
189.24 (0.97)
NA

Values in the bracket are the correlation coefcient. NA, not available.

50

Y. Zhao et al. / Thermochimica Acta 555 (2013) 4652

Fig. 6. FTIR spectra of the commercial PF resin at different degradation temperatures.

Fig. 7. FTIR spectra of the lab PF resin at different degradation temperatures.

groups of the tested resins changed with the increase of thermal


degradation temperature.
The intensity of the OH peak at around 33503500 cm1 for
all the tested resins decreased when temperature was increased
from room temperature to 400 C. This was due to the condensation process involving either residual methylol groups or the
phenolic hydroxyl groups and the formation of the new crosslinks. Possible reactions are shown in Fig. 8 [4,5]. The OH peak
intensity at 1353 cm1 also decreased with the increase of temperature due to the same mechanisms. When the temperature
was further increased to above 600 C, the intensity of the OH
peak at 33503500 cm1 decreased signicantly and the peak at
1353 cm1 disappeared. The main structures of the resins changed
to poly-aromatic structures [4]. The intensity of the peaks at
2914 cm1 and 2848 cm1 due to aliphatic CH2 asymmetric stretch

and aliphatic CH2 symmetric stretch, respectively, decreased with


the increase of thermal degradation temperature.
The peak at 1740 cm1 due to the carboxylic groups was
observed in the bark extractive-PF resin, liqueed bark-PF resin and
lab PF resin when the thermal degradation temperature reached
to 400 C. It also appeared in all the tested resins when the thermal degradation increased to 800 C. The intensity of the peak
was weak, but its appearance was considered as an evidence for
oxidation during the resin thermal degradation under the inert
atmosphere [4,5,18]. The possible reactions are shown in Fig. 9.
Carbonyl groups appearing at 1650 cm1 as a shoulder peak next
to 1600 cm1 were be observed in all the resins when the thermal
degradation temperature reached to 200 C. The intensity of this
peak decreased when the temperature increased. The intensity of
the peak of benzene rings at 1467 cm1 increased for all the resins

Fig. 8. Possible condensation reactions during the thermal degradation process.

Y. Zhao et al. / Thermochimica Acta 555 (2013) 4652

OH

CH2

OH

OH

H2C

+OH

51

OH

HOH2C

+OH
-H2O

OHC

OH

+OH

OH

HOOC

OH

OH

OH

CH2

OH

CH

OH

CO2

O
C

OH

OH

Fig. 9. Possible oxidation reactions during the resin thermal degradation process.

during the thermal degradation process. The intensity of the peak at


879 cm1 assigned to the poly-substituted aromatic ring increased
with the increasing thermal degradation temperature, indicating
the occurrence of the carbonization process. The intensity of the
peak assigned to the methylene bridges at 1435 cm1 decreased
when the temperature increased, reecting the occurrence of the
resin network decomposition.
The peak of alkyl-phenol C O stretching at 1213 cm1 and
1170 cm1 decreased with the increasing thermal degradation
temperature. It disappeared when the temperature was above
600 C for lab PF, commercial PF and bark extractive-PF resins. The
alkyl-phenol C O stretching peak of the liqueed bark-PF resin was
still present even when the temperature reached 800 C.
For the bark extractive-PF resin, the peak at 1249 cm1
decreased at 200 C but increased at 400 C, indicating formation
of ether linkage. While for the lab PF and the commercial PF
resins, the peak at 1249 cm1 was absent in the initial spectra
but appeared when the thermal degradation temperature reached
400 C. It indicated that the oxidation occurred during the resin
thermal degradation process.
When the thermal degradation temperature was lower than
400 C, the polymer network of the phenolic resins remains essentially intact. When the thermal degradation temperature was above
600 C, dramatic changes were seen in the FTIR spectra of all the
resins due to the collapse of the polymeric network to form polyaromatics, which was consistent with past studies on PF resins
[4,18].
It is also noteworthy that the differences between the FTIR spectra of the commercial PF resins after heating at 200 C and 400 C.
According to a previous study [25], the peaks at 2227 cm1 and
2167 cm1 attributed to CN triple bond in cyanamide and the NCO
groups were observed when the thermal degradation temperature
was 200 C. The intensity of the peak at 2167 cm1 decreased significantly while the intensity of the peak at 2227 cm1 increased when

the temperature increased to 400 C. Both peaks disappeared when


the thermal degradation temperature increased to 600 C. The reason for the change may be due to the thermal degradation of urea in
the commercial PF resin. The possible reactions that were reported
before are shown in Fig. 10.

4. Conclusions
The post-curing thermal stability of the bark-derived PF resins
was found to be similar to that of the lab PF resin but differed signicantly from that of the commercial PF resin. The bark
extractive-PF resin and the phenol-liqueed bark-PF resin exhibited different thermal stability and thermal degradation kinetics.
The bark extractive-PF resin showed a better thermal stability. The
structural changes and the differences in structures among the
bark-derived PF resins, the lab PF resin and the commercial PF resin
during the thermal degradation process were revealed by the FTIR
analysis. The commercial PF resin showed different FTIR spectral
characteristics from those of the bark-derived PF resins and lab PF
resin when the thermal degradation temperature was at 200 C and
400 C. The thermal degradation mechanism involving oxidations
was observed for the bark-derived PF resins.

Acknowledgements
Financial support from the Ontario Research Fund-Research
Excellence: Bark Biorenery program partners is greatly appreciated. Authors would also like to thank FPInnovations for providing
the bark samples and the commercial PF resin. We are deeply grateful to Dr. Syed Abthagir Pitchai Mydeen for helping with the TGA
measurements.

References

Fig. 10. Possible reactions for urea decomposition.

[1] A. Pizzi, Wood Adhesives Chemistry and Technology, vol. 1, Marcel Dekker,
New York, 1993.
[2] V.A. Er, A. Mattila, Thermal analysis of thermosetting resins, J. Therm. Anal. 10
(1976) 461469.
[3] T.R. Manley, Thermal analysis of polymers, Pure Appl. Chem. 61 (1989)
13531360.
[4] Y.F. Chen, Z.B. Chen, S.Y. Xiao, H.B. Liu, A novel thermal degradation mechanism
of phenolformaldehyde type resins, Thermochim. Acta 47 (2008) 639643.

52

Y. Zhao et al. / Thermochimica Acta 555 (2013) 4652

[5] L.L. Costa, R.D. Montelera, G. Camino, E.D. Weil, E.M. Pearce, Structurecharring
relationship in phenol formaldehyde type resins, Polym. Degrad. Stab. 56
(1997) 2335.
[6] M.S. Chetan, R.S. Ghadage, C.R. Rajan, V.G. Gunjikar, S. Ponrathnam, Thermolysis of orthonovolacs. Part 1. Phenolformaldehyde and m-cresol formaldehyde
resins, Thermochim. Acta 228 (1993) 261270.
[7] M.S. Chetan, R.S. Ghadage, C.R. Rajan, V.G. Gunjikar, S. Ponrathnam, Thermolysis
of orthonovolacs. Part 2. Phenolformaldehyde and -naphtholformaldehyde
resins, J. Appl. Polym. Sci. 50 (1993) 685692.
[8] G.P. Shulman, H.W. Lochte, Thermal degradation of polymers. 2. Mass spectrometric thermal analysis of phenolformaldehyde polycondensates, J. Appl.
Polym. Sci. 10 (1996) 619635.
[9] S. Vyazovkin, A.K. Burnham, J.M. Criado, L.A. Prez-Maqueda, C. Popescu,
N. Sbirrazzuoli, ICTAC Kinetics Committee recommendations for performing
kinetic computations on thermal analysis data, Thermochim. Acta 520 (2011)
119.
[10] C. Popescu, Integral method to analyze the kinetics of heterogeneous reactions
under non-isothermal conditions A variant on the OzawaFlynnWall method,
Thermochim. Acta 285 (1996) 309323.
[11] J.H. Flynn, L.A. Wall, General treatment of the thermogravimetry of polymers,
J. Res. NBS: A Phys. Chem. 70 (1966) 487523.
[12] T. Ozawa, A new method of analyzing thermogravimetric data, Bull. Chem. Soc.
Jpn. 38 (1965) 18811886.
[13] S. Vyazovkin, Modication of the integral isoconversional method to
account for variation in the activation energy, J. Comput. Chem. 22 (2001)
178183.
[14] J.J.M. rfo, F.G. Martins, Kinetic analysis of thermogravimetric data obtained
under linear temperature programming a method based on calculations
of the temperature integral by interpolation, Thermochim. Acta 390 (2002)
195211.
[15] G. Gyulai, E.J. Greenhow, A new integral method for the kinetic analysis of
thermogrametric data, J. Therm. Anal. 6 (1974) 279291.
[16] M.R. Rao, S. Alwan, K.J. Scariah, K.S. Sastri, Thermochemical characterization of
phenolic resins thermogravimetric and pyrolysis-GC studies, J. Therm. Anal. 49
(1997) 261268.

[17] M.R. Rao, B.S.M. Rao, C.R. Rajan, R.S. Ghadage, Thermal degradation kinetics of
phenolcrotonaldehyde resins, Polym. Degrad. Stab. 61 (1998) 283288.
[18] M.V. Alonso, M. Oliet, J.C. Domnguez, E. Rojo, F. Rodrguez, Thermal degradation of ligninphenolformaldehyde and phenolformaldehyde resol resins.
Structural changes, thermal stability, and kinetics, J. Therm. Anal. Calorim. 105
(2011) 349356.
[19] Y.K. Lee, D.J. Kim, H.J. Kim, T.S. Hwang, M. Rafailovich, J. Sokolov, Activation
energy and curing behavior of resol- and novolac-type phenolic resins by differential scanning calorimetry and thermogravimetric analysis, J. Appl. Polym.
Sci. 89 (2003) 25892596.
[20] J.M. Prez, A. Fernndez, Thermal stability and pyrolysis kinetics
of ligninphenolformaldehyde resins, J. Appl. Polym. Sci. (2011),
http://dx.doi.org/10.1002/app.34817.
[21] C.N. Zrate, M.I. Aranguren, M.M. Reboredo, Thermal degradation of a phenolic
resin, vegetable bers, and derived composites, J. Appl. Polym. Sci. 107 (2008)
29772985.
[22] Y. Zhao, N. Yan, M. Feng, Characterization of phenolformaldehyde resins
derived from liqueed lodgepole pine barks, Int. J. Adhes. Adhes. 30 (2010)
689695.
[23] Y. Zhao, N. Yan, M. Feng, Characterization of phenolformaldehyde resins
derived from lodgepole pine bark extractives, J. Adhes. Sci. Technol. (2012),
http://dx.doi.org/10.1080/01694243.2012.697689.
[24] F. Yao, Q.L. Wu, Y. Lei, W.H. Guo, Y.J. Xu, Thermal decomposition kinetics of natural bers: activation energy with dynamic thermogravimetric analysis, Polym.
Degrad. Stab. 93 (2008) 9098.
[25] P.M. Schaber, J. Colson, S. Higgins, D. Thielen, B. Anspach, J. Brauer, Thermal
decomposition (pyrolysis) of urea in an open reaction vessel, Thermochim. Acta
424 (2004) 131142.
[26] I. Poljansek, M. Krajnc, Characterization of phenolformaldehyde prepolymer
resins by in line FT-IR spectroscopy, Acta Chimi. Slovenia 52 (2005) 238244.
[27] Y. Zhao, N. Yan, M. Feng, Bio-based phenol formaldehyde resins derived from
beetle infested pine barks-structure and composition, ACS Sust. Chem. Eng.
(2013), http://dx.doi.org/10.1021/sc3000459.
[28] Y. Zhao, N. Yan, M. Feng, Effects of reaction conditions on phenol liquefaction
of beetle infested lodgepole pine barks, Curr. Org. Chem. (2013).

You might also like