Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Industrial Crops and Products 49 (2013) 412–418

Contents lists available at SciVerse ScienceDirect

Industrial Crops and Products


journal homepage: www.elsevier.com/locate/indcrop

Use of tung oil as a reactive toughening agent in


dicyclopentadiene-terminated unsaturated polyester resins
Chengguo Liu a,b,∗ , Wen Lei c , Zhengchun Cai c , Jianqiang Chen c , Lihong Hu a,b ,
Yan Dai a , Yonghong Zhou a,∗∗
a
Institute of Chemical Industry of Forestry Products, Chinese Academy of Forest, Nanjing 210042, PR China
b
Institute of Forest New Technology, CAF, Beijing 100091, PR China
c
College of Science, Nanjing Forestry University, Nanjing 210037, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Unsaturated polyester resin terminated with dicyclopentadiene (DCPD-UPR) was modified by tung oil
Received 25 March 2013 (TO) via intermolecular Diels–Alder reaction occurring at the later stage of melt polycondensation. These
Received in revised form 23 May 2013 TO-modified DCPD-UPR (DCPD-UPR-TO) polymers were characterized by FT-IR, 1 H NMR, and gel per-
Accepted 24 May 2013
meation chromatography. The DCPD-UPR-TO polymers were further blended with styrene comonomer
and cured via free-radical polymerization to give crosslinked thermosetting polymers. Scanning elec-
Keywords:
tron microscopy was employed to produce surface morphology and dynamic mechanical analysis was
Tung oil
conducted to study the thermo-mechanical properties. Thermal and mechanical properties of these bio-
Toughening agent
Unsaturated polyester resin
materials were also investigated and the results show that toughness of them increases obviously as the
Dicyclopentadiene increase of TO content. Compared with the neat DCPD-UPR matrix, the matrix obtained from DCPD-UPR-
Diels–Alder reaction TO with a TO content of 20% has maximum increase of 373% and 875% in impact strength and tensile
failure strain due to the synergistic effects of phase separation and crosslink density. However, the opti-
mum amount of tung oil is 10% because its polymer matrix gives a best stiffness–toughness balance. These
oil-based polymer materials show promise as an alternative to replace petroleum-based materials.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction high-added-value areas. As a result, the flexibility and toughness of


DCPD-UPRs need to be improved.
Unsaturated polyester resins (UPRs) are currently the most In the past several decades, numerous soft materials have been
widely utilized thermosetting polymer because of its low cost, used to address this obstacle for general UPRs. For example, there
ease of handling, and good balance of mechanical, electrical, and are several reports on the modification of UPRs by blending with
chemical properties. The reinforced composite materials and non- rubbers (Pachpinyo et al., 2006; Ray, 2009), diphenylmethane-4,4 -
reinforced materials of UPRs have been broadly employed in diisocyanate (Gunduz et al., 2005) or polyurethane prepolymers
aerospace, automotive, marine, infra-structure, military, sports, having terminal isocyanate groups (Cherian et al., 2006), etc. Due
industrial fields, etc. (Mighani, 2012). Dicyclopentadiene (DCPD), a to the threats of uncertain petroleum supply in the future and envi-
byproduct from the pyrolysis of petroleum for C5 compounds, has ronmental pollution, there is a growing trend to incorporate natural
been employed to synthesize DCPD-modified UPRs (DCPD-UPRs) oils (renewable resources), especially their derivates, into UPRs as
since 1980s (Krupka et al., 2008; Pan et al., 2012). The excellent impact modifiers. Mehta et al. (2004) employed methyl ester of
stiffness, corrosion resistance, and air-drying property, combin- soybean oil and epoxidized methyl linseedate (EML) to toughen
ing with its economical price, make DCPD-UPRs important in UPR UPR, and prepared biocomposites from the modified UPRs. The
products. However, the brittle nature of them at ambient temper- notched Izod impact strength of biocomposites from the biobased
ature limits their application in fiber-reinforced plastics and other UPR blends and hemp fiber mat are enhanced by 90% by comparison
with that of the pure UPR/hemp fiber mat composites. Miyagawa
et al. (2006, 2007) used UPRs containing epoxidized methyl soyate
(EMS) or EML to prepare novel biobased resins. The resulting new
∗ Corresponding author at: Institute of Chemical Industry of Forestry Products, biobased thermosets showed relatively high Izod impact strength.
Chinese Academy of Forest, Nanjing 210042, PR China. Tel.: +86 025 85482520; Haq et al. (2009, 2011) prepared biobased resins by a partial substi-
fax: +86 025 85482777.
∗∗ Corresponding author. Tel.: +86 025 85482777; fax: +86 025 85482777. tution of UPR with EMS or EML, and reinforced them with natural
E-mail addresses: liuchengguo@gmail.com (C. Liu), yhzhou777@sina.com fibers, nanoclays or layered silicates. Toughness of the biobased
(Y. Zhou). resins was improved by the corporation of EMS or EML. Ghorui et al.

0926-6690/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.indcrop.2013.05.023
C. Liu et al. / Industrial Crops and Products 49 (2013) 412–418 413

(2011) used maleated castor oil (MACO) as biomodifier in UPR/fly to a set value. The obtained DCPD-UPR polymers were a light trans-
ash composites. By the incorporation of 5 wt% MACO, the impact parent liquid at 90 ◦ C.
strength of the UPR matrix increased by 52% without any loss in
modulus. Das et al. (2011) prepared novel biomaterials through 2.3. Synthesis of TO-modified DCPD-UPR
the blending of an unsaturated polyester resin/styrene mix with
tung oil (TO), which offered improved impact strength at a very The synthesis of DCPD-UPR-TO polymers was carried out in
low content of TO. three basic steps, as shown in Fig. 1. The first step is the syn-
In this study, unmodified TO is utilized to improve the tough- thesis of DCPD-MA, which is the same as that in the synthesis of
ness of DCPD-UPRs. Although Das et al. (2011) has reported the DCPD-UPR. The second and third steps were similar to the second
physically blending approach to prepare UPR/TO materials, several step shown in Section 2.2. At the latter stage of melt polyconden-
distinct features are demonstrated in this work. First, TO is incor- sation, when the acid value of the reacting mixture decreased to
porated into UPR via the intermolecular Diels–Alder (D–A) reaction 60–70 mgKOH/g, the reactor was cooled to about 160 ◦ C and a cer-
between double (C C) bonds on polyester chains and TO conju- tain amount (for example, 16.7 g) of tung oil was added in drops.
gated triene groups (Fig. 1). Second, the D–A reaction occurs at the After this the reactor was heated to 200 ◦ C again and maintained at
later stage of melt polycondensation for polyesters. Third, there is this temperature until the acid value decreased to the set value. The
limited report on the modification of DCPD-UPR by natural oils. The obtained DCPD-UPR-TO polymers were light yellow opaque liquids
TO-modified DCPD-UPR (DCPD-UPR-TO) polymers are character- at 90 ◦ C.
ized by FT-IR, 1 H NMR, and gel permeation chromatography (GPC). In our experiments, the contents of TO in the obtained
The structures and properties of the cured DCPD-UPR-TO polymer DCPD-UPR-TO polymers were 5%, 10%, 15%, and 20%, thus the corre-
matrices are also investigated. sponding DCPD-UPR-TO polymers were labeled as DCPD-UPR-TO5,
DCPD-UPR-TO10, DCPD-UPR-TO15, and DCPD-UPR-TO20, respec-
tively. The neat DCPD-UPR can be regarded as the DCPD-UPR-TO
2. Materials and methods polymer with 0% content of tung oil.

2.1. Materials 2.4. Curing of DCPD-UPR-TO polymers with styrene

Tung oil was a commercial product purchased from Jiangsu All the obtained neat DCPD-UPR and DCPD-UPR-TO poly-
Donghu Bio-energy Plant Plantation (China), which has a yellow mer products were blended with styrene (64.2 g, 0.62 mol) and
color and a specific gravity of 0.935–0.940 at 25 ◦ C. Diethyl- hydroquinone (0.02 g) at 90 ◦ C for 1 h to give liquid resins. The
ene glycol (DEG), 1,2-propanediol (PG), maleic anhydride (MA), as-prepared DCPD-UPR-TO/styrene resins were cured following a
phthalic anhydride (PA), ethanol, potassium hydroxide (KOH), and same procedure: blended with the initiator (2 wt% of the resins)
phosphoric acid (H3 PO4 ) were obtained from Shanghai Lingfeng for 20 min and then with the promoter (0.2 wt% of the resins) for
Chemical Reagent Co. Ltd. (China). Dicyclopentadiene (DCPD) with 1–2 min, poured into molds, cured at room temperature for 24 h,
a purity of 90% was obtained from Hangzhou Yangli Shihua Co. Ltd. and postcured at 60 ◦ C for 3 h. After that the neat DCPD-UPR and
(China). Hydroquinone (≥98%) and styrene (≥98%) were obtained DCPD-UPR-TO polymer matrices were acquired, which could be
from Chengdu Kelong Chemical Reagent Co. Ltd. (China). The initia- used for the following characterization of structure and property.
tor benzoyl peroxide (≥98%) and the promoter N,N-dimethylaniline
(≥98%) were obtained from Shanghai Aladdin Chemistry Co. Ltd. 2.5. Characterization
(China). The DEG, PG, and styrene were dried by molecular sieves
at least for 1–2 weeks before use. All the other reagents were used Reaction procedure of melt polycondensation was monitored by
as received. measuring acid values with 1 h time interval, using KOH/ethanol as
standard solution. The acid value was calculated as (Ghorui et al.,
2011)
2.2. Synthesis of DCPD-UPR
56.1 · V · CKOH
Av = (1)
The synthesis of DCPD-UPR was performed in two basic steps, as 1000 · m
partially shown in Fig. 1. First, dicyclopentadiene maleate (DCPD- where V, CKOH represents consumed volume and initial concen-
MA) was synthesized via hydrolysis: a reaction mixture of MA tration of KOH solution, respectively; m is the sample weight.
(39.2 g, 0.4 mol), DEG (12.7 g, 0.12 mol), deionized water (7.5 g, Viscosity (Vs ) measurements for the DCPD-UPR-TO/styrene
0.42 mol), and H3 PO4 (0.3 g), was added into a 250-mL four-neck resins were performed on a NDJ-8S rotational viscometer (Shanghai
flask; the flask was then equipped with mechanical stirrer, ther- Changji Dizhi Instrument Corporation, China). The IR spectra
mometer, nitrogen (N2 ) gas inlet, and reflux condensation tube. were recorded on a Nicolet iS10 IR spectrometer (Thermo-Fisher
The reactor was subsequently heated by oil bath to 70 ◦ C and agi- Corporation, USA). The 1 H NMR spectra were recorded on a
tated at this temperature for 30 min until the MA solid completely Bruker DRX-300 Advance NMR spectrometer (Bruker Corpora-
melt. Then it was heated to 125–135 ◦ C under N2 atmosphere, and tion, Germany). The GPC chromatographs were recorded on a
DCPD liquid (41.1 g, 0.28 mol) was added drop by drop. After this GPC instrument with Styragel HR5E and HR2 (300 mm × 7.8 mm,
the reacting mixture maintained at this temperature for 2 h to give Waters Corporation, USA) columns as well as equipped with an
a light and transparent liquid. Second, DCPD-modified unsaturated Optilab-rEX refractive index detector (Wyatt Technology Corpora-
polyester resin (DCPD-UPR) was prepared via melt polyconden- tion, USA). HPLC-grade tetrahydrofuran (THF) was used as eluent
sation: a mixture of DEG (5.3 g, 0.05 mol), PG (25.1 g, 0.33 mol), and the flow rate of the eluent was 1.0 mL/min.
PA (29.6 g, 0.2 mol), and hydroquinone (0.02 g) was consecutively Tensile and flexural tests of the polymer matrices were evalu-
added into the reactor. The reflux condensation device was changed ated using a SANS7 CMT-4304 universal tester (Shenzhen Xinsansi
into to a fractional distillation device. The precondensation was Jiliang Instrument Corporation, China). Impact tests of the sam-
conducted under N2 atmosphere at 160 ◦ C for 1.5 h. After gradually ples were performed in a XJJY-5 impact tester (Chengde Xinguo
heating to 200 ◦ C, the polycondensation lasted for 5 h at this tem- Instrument Corporation, China). All the above tests followed the
perature until the acid value (Av ) of the reacting mixture decreased procedure specified in the GB/T 2567-2008 Standard. For accuracy
414 C. Liu et al. / Industrial Crops and Products 49 (2013) 412–418

Fig. 1. Reaction scheme for the modification of DCPD-UPR by TO triglyceride molecule.

at least five specimens were measured and the values were UPRs, is very weak for the DCPD-UPR, which is attributed to the
averaged. Thermogravimetric analysis (TGA) was performed on a termination reaction between DCPD and the hydroxyls on UPR.
STA 409PC thermogravimetry instrument (Netzsch Corporation, After the incorporation of TO into DCPD-UPR, two obvious varia-
Germany) at a heating rate of 15 ◦ C/min. Dynamic mechanical anal- tions have been found in the spectrum of DCPD-UPR-TO5. First, the
ysis (DMA) was conducted in three-point bending geometry on characteristic features of the TO conjugated triene at 3012, 991,
a Q800 solids analyzer (TA Corporation, USA). Scanning electron and 964 cm−1 are not observed. Second, the intensity of the peaks
microscopy (SEM) examinations of the impact-fractured samples at around 2963 cm−1 greatly increases relative to that of the C O
were performed on an S-3400N Scanning Electron Microscope peak at around 1730 cm−1 . These changes indicate that the con-
(HITACHI Corporation, Japan). The surface of the fractured samples jugated triene on TO molecules has reacted with the unsaturated
after completion of the impact tests was coated with a gold film double bonds on DCPD-UPR chains.
prior to SEM observation. 1 H NMR technique is another important tool employed to inves-

tigate the structure of the obtained DCPD-UPR-TO polymers. The


spectra of DCPD-UPR and DCPD-UPR-TO5 are depicted in Fig. 3.
3. Results and discussion The double peak at 7.5–8.0 ppm is corresponding to the protons on
the benzene structure from phthalic anhydride. It can be taken as a
3.1. Structure of the TO-modified DCPD-UPR polymers

The results of ultimate acid value (Av ), viscosity (Vs ), weight- and
number-average molar masses (Mw and Mn ) as well as polydisper-
sity (PDI) for DCPD-UPR-TO polymers with different TO contents
are shown in Table 1. The Av values are very low and almost the
same (about 20 mgKOH/g), indicating the completion of melt poly-
condensation for all the DCPD-UPR-TO polymers. As the increase of
TO contents, viscosities for the DCPD-UPR-TO/styrene resins grad-
ually increase and reach a maximum value of about 1370 mPa s. The
UPRs with these Vs values are in a range suitable for the application
in fiber reinforced materials.
FT-IR technique was employed first to study the structure of
TO-modified DCPD-UPRs. The FT-IR spectra of DCPD-UPR, TO, and
DCPD-UPR-TO5 are depicted in Fig. 2. In the spectrum of TO, there
are several characteristic peaks: the conjugated triene structure
(3012, 991, and 964 cm−1 ), methyl and methylene groups (2925
and 2854 cm−1 ), and ester carbonyl groups (1744 cm−1 ) (Liu et al.,
2012). In the spectrum of DCPD-UPR, there are also some typical
peaks: methyl and methylene groups (2963 and 2887 cm−1 ), C O
on ester group (1730 cm−1 , strong), and C O (1000–1300 cm−1 ).
The hydroxyl peak at 3000–3500 cm−1 , usually strong for general Fig. 2. FT-IR spectra of (a) DCPD-UPR, (b) TO, and (c) DCPD-UPR-TO5.
C. Liu et al. / Industrial Crops and Products 49 (2013) 412–418 415

Fig. 4. GPC elution chromatograms of the DCPD-UPR-TO polymers with different


contents of TO.

indicate the occurrence of the D–A reaction between DCPD-UPR and


TO. Moreover, it is found that the DCPD-UPR-TO polymer almost
reach maximum values of Mw and Mn at a TO content of 10%, after
1
Fig. 3. H NMR spectra of (a) DCPD-UPR and (b) DCPD-UPR-TO5. that the molar masses vary slightly. The reason for this may lie in
that the stoichiometric equilibrium of DCPD-UPR and TO in the D–A
reference, because the intensity of this peak should not be altered reaction is at a ratio of about 9:1 in weight. In other words, it means
throughout the synthesis process of DCPD-UPR-TO polymers. The that TO molecules can completely react with DCPD-UPR chains until
peaks at 6.3 ppm and 6.9 ppm represent the maleate and fumarate the TO content is up to a value of about 10%. When adding more TO
vinyl protons from maleic anhydride, respectively. It is found that (above 10%) into DCPD-UPR, redundant tung oil molecules are left
the ratios of fumarate to maleate vinyl groups are 4.13 and 1.67 for in DCPD-UPR-TO polymers, thus leading to smaller molar masses
DCPD-UPR and DCPD-UPR-TO5 polymers, respectively, indicating (especially Mn ) and larger PDI values.
the occurrence of maleate to fumarate (cis to trans) isomerization.
Compared the spectrum of DCPD-UPR with that of DCPD-UPR- 3.2. Structures and properties of the cured DCPD-UPR-TO
TO5 (Fig. 3(a)), the maleate vinyl protons on the DCPD-UPR-TO5 matrices
(Fig. 3(b)) almost maintain constant, while the fumarate vinyl pro-
tons have a great decrease from 0.62 to 0.30. Two important facts The failure surface morphologies of the matrices obtained from
are indicated here: first, the TO conjugated triene only reacts with DCPD-UPR-TO polymers with different TO contents were investi-
the fumarate structure on DCPD-UPRs in the D–A reaction; second, gated by SEM. Fig. 5 demonstrates the failure-surface morphologies
about 50% of the fumarate vinyl groups have been consumed by the after the completion of impact tests. In Fig. 5(a), the failure surface
D–A reaction. In addition, the peaks at 3.0–3.3 ppm demonstrate the of the DCPD-UPR matrix is generally flat and featureless. This sug-
protons at the place where the unsaturated double bond is linking gests that the behavior of the matrix is linear elastic, and the crack
to the conjugated triene group (Liu et al., 2012). The appearance of propagates in a planar manner under the impact loading. However,
the peak at 0.9 ppm belongs to the terminal methyl protons of TO the failure surface of the TO-modified DCPD-UPR matrices become
fatty acids ( CH3 ). rougher, which is mainly due to the soft structure provided by the
GPC technique is employed to analyze molar masses of the linked or the unreacted TO triglycerides. It was reported that the
obtained DCPD-UPR-TO polymers. The elution chromatographs of rougher surface was identical for dissipating more energy that was
GPC are shown in Fig. 4 and the corresponding results of molar due to shear deformation during the crack propagation (Miyagawa
masses are listed in Table 1. As shown in Fig. 4, the area of elu- et al., 2005). Thus it is expected that the increased roughness here
tion time at 11–13 min, which corresponds to the section of high will lead to a pronounced toughening effect on the brittle DCPD-
molar mass, has greatly increased for DCPD-UPR-TO polymers by UPR matrix. In addition, it is found that the DCPD-UPR-TO matrices
the incorporation of TO molecules. This fact is further evidenced by with TO contents from 0 to 10% do not show phase separation, while
the change of molar masses for the obtained DCPD-UPR and DCPD- the matrices with TO contents above 10% demonstrate obvious
UPR-TO polymers (Table 1). It is observed that both the weight- phase separation such as small resin pieces. This phenomenon also
and number-average molar masses (Mw and Mn ) of DCPD-UPR- indicates that TO can completely react with DCPD-UPR chains up
TO polymers are larger than those of neat DCPD-UPR. All of these to a TO content of about 10%, which is in good agreement with the

Table 1
Acid value (Av ), viscosity (Vs ), weight- and number-average molar masses (Mw and Mn ) as well as polydispersity (PDI) for DCPD-UPR-TO polymers with different TO contents.

Samples Av (mgKOH/g) Vs (mPa s) Mw (g/mol) Mn (g/mol) PDI

DCPD-UPR 19.9 840 1275 618 2.06


DCPD-UPR-TO5 16.3 967 2103 737 2.85
DCPD-UPR-TO10 20.3 1370 3145 858 3.67
DCPD-UPR-TO15 22.1 1185 2993 807 3.71
DCPD-UPR-TO20 23.6 1348 3170 818 3.88
416 C. Liu et al. / Industrial Crops and Products 49 (2013) 412–418

Fig. 5. SEM micrographs of the impact fracture surfaces of the polymer matrices obtained from DCPD-UPR-TO polymers with different TO contents.

GPC analysis. This phase separation phenomenon is also expected namely the average number of crosslinks per unit volume, can be
to affect the following properties for the TO-modified DCPD-UPR determined from the rubbery modulus using the following equa-
biomaterials. It is noteworthy that the true TO weight fraction tion (Lu et al., 2005; Miyagawa et al., 2005; Can et al., 2006; Liu
in the cured DCPD-UPR-TO10 matrix is about 7.4%. This value is et al., 2013)
much higher than that reported by Das et al. (2011), in whose
E  = 3e RT (2)
work through a simple physically blending approach the prepared
UPR/TO matrix showed clear phase separation at a TO content of where E represents the storage modulus of the crosslinked copoly-
about 2%. mer in the rubbery plateau region, R is the gas constant, and T is the
DMA was conducted to study thermo-mechanical properties absolute temperature. As shown in Table 2, the crosslink densities
of the obtained bioplastics. The results are depicted in Fig. 6 and of all the matrices are in the range of 2810–2290 mol/m3 , which
Table 2. It is seen that the storage moduli of all the DCPD-UPR-TO is corresponding to crosslink molar masses of 391–480 g/mol. It is
bioplastics at 25 ◦ C are lower than that of the neat DCPD-UPR, which found that both the storage moduli at 25 ◦ C and 120 ◦ C as well as
means the dynamic mechanical properties of DCPD-UPR-TO matri- the crosslink density decrease as the increase of TO content. This is
ces decrease by the addition of TO. This fact is also demonstrated by mainly because the intermolecular D–A reaction between polyester
the storage modulus in the rubbery region (at 120 ◦ C in this study), and TO reduces the amount of crosslinkable C C bonds in the
as shown in Table 2. Based on the kinetic theory of rubber elastic- polyester chains. Glass transition temperature (Tg ) is determined
ity, the experimental crosslink density of all the copolymers (ve ), from the peak of loss factor (tan ı). It can be seen from Fig. 6(b)

Fig. 6. Dynamic mechanical analysis in (a) storage modulus and (b) damping parameter for the polymer matrices obtained from DCPD-UPR-TO polymers with different TO
contents.
C. Liu et al. / Industrial Crops and Products 49 (2013) 412–418 417

Table 2
Storage modulus (E ), glass transition temperature (Tg ), crosslink density (e ), effective molar mass between crosslinks (Mc ), 10% and 50% weight loss temperatures (T10 and
T50 ) for neat DCPD-UPR and DCPD-UPR-TO polymer matrices.

Samples E at 25 ◦ C (GPa) E at 120 ◦ C (MPa) Tg (◦ C) e (103 mol/m3 ) Mc (g/mol) T10 (◦ C) T50 (◦ C)

DCPD-UPR 1.53 27.5 79.3 2.81 391 294.5 382.0


DCPD-UPR-TO5 1.26 26.8 61.8 2.73 403 294.5 384.5
DCPD-UPR-TO10 1.24 25.4 65.6 2.59 424 302.0 389.5
DCPD-UPR-TO15 0.802 23.6 60.7 2.40 458 294.5 387.0
DCPD-UPR-TO20 0.490 22.4 53.1 2.29 480 292.0 389.5

unreacted oil or other soluble components in the bulk material. The


stage 2 (about 320–460 ◦ C) is the fastest and corresponds to degra-
dation and char formation of the crosslinked polymer structure,
while the last stage (>460 ◦ C) corresponds to gradual degradation
of the char residue. Table 2 summarizes the thermal data of these
biobased materials, including 10% and 50% mass loss temperatures
(T10 and T50 ). It is clearly seen that both of the two temperatures
increase to maximum values at the TO content of 10%, which is
mainly caused by the phase separation effect as mentioned above.
Compared with the neat DCPD-UPR matrix, the DCPD-UPR-TO10
matrix has a maximum increase of 7.5 ◦ C in both T10 and T50 . Hence,
the DCPD-UPR-TO10 matrix shows slightly better thermal stability
than the pure DCPD-UPR matrix.
The changes in tensile strength and modulus, flexural strength
and modulus of the matrices obtained from DCPD-UPR-TO poly-
mers with different TO contents are shown in Fig. 8. It is observed
that all these properties decrease as the increase of TO content,
Fig. 7. TGA curves of the polymer matrices obtained from DCPD-UPR-TO polymers indicating the decrease of stiffness for all the DCPD-UPR-TO poly-
with different TO contents. mer matrices. This is mainly because of the decrease of crosslink
density, which was reported that its decrease resulted in the stiff-
and Table 2 that Tg for the DCPD-UPR-TO bioplastics drops by the ness decrease of a polymer matrix (Mehta et al., 2004; Miyagawa
addition of TO and has a minimum decrease of 13.7 ◦ C for the DCPD- et al., 2005; Haq et al., 2011; Liu et al., 2013). Interestingly, the
UPR-TO10 polymer matrix by comparison with that of the neat decrease in stiffness seems to have two distinct regions: at a
DCPD-UPR matrix. It may be caused by a combined effect of phase TO content below 10%, the stiffness of DCPD-UPR-TO biomateri-
separation and crosslink density, because the decrease of crosslink als gradually decreases; at a TO content above 10%, the stiffness
density will result in the decrease of Tg (Can et al., 2006) and the decreases rapidly. This phenomenon can be attributed to the phase
boundary of phase separation phenomenon at the TO content of separation effect, as shown in the above SEM images. As mentioned
10% make the DCPD-UPR-TO10 matrix less influenced. in previous works (Mehta et al., 2004; Miyagawa et al., 2005; Ghorui
Fig. 7 demonstrates TGA thermograms of the DCPD-UPR-TO et al., 2011; Haq et al., 2011; Liu et al., 2013), the introduction of
polymer matrices with different TO contents. It is observed that a second rubber phase like oil-rich phase would sacrifice the stiff-
all the bioplastics are thermally stable in nitrogen gas below 180 ◦ C ness of a polymer matrix but improve its toughness. Therefore, it
and exhibit a three-stage thermal degradation above this tempera- is concluded that the phase separation occurring at the TO content
ture (Andjelkovic et al., 2005). The first stage degradation (about above 10% results in the rapid decrease of stiffness for the DCPD-
180–320 ◦ C) is attributed to evaporation and decomposition of UPR-TO matrices. Compared with the neat DCPD-UPR matrix, the

Fig. 8. Tensile and flexural properties in (a) strength and (b) modulus of the polymer matrices obtained from DCPD-UPR-TO polymers with different TO contents.
418 C. Liu et al. / Industrial Crops and Products 49 (2013) 412–418

benefit to the modification of petroleum-based materials with soft


bioresources. These biomaterials including renewable resources
(e.g. tung oil) are very promising in application of fiber-reinforced
composite materials.

Acknowledgements

The authors are grateful to the financial support from Chi-


nese National Special Fund of International Cooperation Of Science
and Technology (No. 2011DFA32440) and Chinese Academy
of Forest for the Special Fund of Fundamental Research (No.
CAFINT2011C02).

References

Andjelkovic, D.D., Valverde, M., Henna, P., Li, F., Larock, R.C., 2005. Novel thermosets
prepared by cationic copolymerization of various vegetable oils—synthesis and
Fig. 9. Impact strength and tensile failure strain of the polymer matrices obtained
their structure–property relationships. Polymer 46, 9674–9685.
from DCPD-UPR-TO polymers with different TO contents. Can, E., Wool, R.P., Küsefoğlu, S., 2006. Soybean- and castor-oil-based thermosetting
polymers: mechanical properties. J. Appl. Polym. Sci. 102, 1497–1504.
Cherian, A.B., Abraham, B.T., Thachil, E.T., 2006. Modification of unsaturated
DCPD-UPR-TO10 matrix demonstrates decrease of 5.3% and 18.8% polyester resin by polyurethane prepolymers. J. Appl. Polym. Sci. 100,
in tensile and flexural strength, 19.3% and 16% in tensile and flexural 449–456.
modulus, respectively, which still can be utilized in fiber reinforced Das, K., Ray, D., Banerjee, C., Bandyopadhyay, N.R., Mohanty, A.K., Misra, M., 2011.
Novel materials from unsaturated polyester resin/styrene/tung oil blends with
materials. high impact strengths and enhanced mechanical properties. J. Appl. Polym. Sci.
Fig. 9 shows the changes in impact strength and tensile failure 119, 2174–2182.
strain of the DCPD-UPR-TO polymer matrices. It is clearly seen that Ghorui, S., Bandyopadhyay, N.R., Ray, D., Sengupta, S., Kar, T., 2011. Use of maleated
castor oil as biomodifier in unsaturated polyester resin/fly ash composites. Ind.
all these properties increase as the increase of TO content, indi- Crop. Prod. 34, 893–899.
cating the increase of toughness for all the DCPD-UPR-TO polymer Gunduz, G., Erol, D., Akkas, N., 2005. Mechanical properties of unsaturated polyester-
matrices. Similarly to the stiffness shown above, the increase in isocyanate hybrid polymer network and its E-glass fiber-reinforced composite.
J. Compos. Mater. 39, 1577–1589.
toughness can be divided into two distinct regions, which can also Haq, M., Burgueno, R., Mohanty, A.K., Misra, M., 2009. Bio-based unsaturated
be attributed to the effects of phase separation and crosslink den- polyester/layered silicate nanocomposites: characterization and thermo-
sity. It is known to us that both the occurrence of phase separation physical properties. Compos. A: Appl. Sci. Manuf. 40, 540–547.
Haq, M., Burgueno, R., Mohanty, A.K., Misra, M., 2011. Bio-based polymer nanocom-
and the decrease of crosslink density will lead to the toughness
posites from UPE/EML blends and nanoclay: development, experimental
improvement of the resulting thermosets (Mehta et al., 2004; characterization and limits to synergistic performance. Compos. A: Appl. Sci.
Ghorui et al., 2011; Haq et al., 2011; Liu et al., 2013). Compared Manuf. 42, 41–49.
with the neat DCPD-UPR matrix, the DCPD-UPR-TO10 matrix shows Krupka, J., Stepanek, K., Herink, T., 2008. Dicyclopentadiene and its derivatives in
chemical industry. Chem. Listy 102, 1107–1114.
increases of 72.5% and 91.3% in impact strength and tensile fail- Liu, C.G., Dai, Y., Wang, C.S., Xie, H.F., Zhou, Y.H., Lin, X.Y., Zhang, L.Y., 2013.
ure strain, while the increases in them for the DCPD-UPR-TO20 Phase-separation dominating mechanical properties of a novel tung-oil-based
matrix are 373% and 875%, respectively. Such large improvement in thermosetting polymer. Ind. Crop. Prod. 43, 677–683.
Liu, C.G., Yang, X.H., Cui, J.F., Zhou, Y.H., Hu, L.H., Zhang, M., Liu, H.J., 2012. Tung oil
toughness is due to the synergistic effects of phase separation and based monomer for thermosetting polymers: synthesis, characterization, and
crosslink density. Hence, the toughness of the DCPD-UPR matrix is copolymerization with styrene. Bioresources 7, 447–463.
greatly improved by the chemical approach we used to incorporate Lu, J., Khot, S., Wool, R.P., 2005. New sheet molding compound resins from soybean
oil. I. Synthesis and characterization. Polymer 46, 71–80.
TO into DCPD-UPR. Mehta, G., Mohanty, A.K., Misra, M., Drzal, L.T., 2004. Biobased resin as a toughening
agent for biocomposites. Green Chem. 6, 254–258.
4. Conclusion Mighani, H., 2012. Synthesis of thermally stable polyesters. In: Saleh, H.E.-D.M. (Ed.),
Polyester. InTech, Rijeka, pp. 3–17.
Miyagawa, H., Misra, M., Drzal, L.T., Mohanty, A.K., 2005. Fracture toughness
In this work novel biobased resins, TO-modified DCPD-UPRs, and impact strength of anhydride-cured biobased epoxy. Polym. Eng. Sci. 45,
have been successfully synthesized via the intermolecular D–A 487–495.
Miyagawa, H., Mohanty, A.K., Burgueno, R., Drzal, L.T., Misra, M., 2006. Development
reaction between double bonds on polyester chains and TO con- of biobased unsaturated polyester containing functionalized linseed oil. Ind. Eng.
jugated triene. The structure characterization of the obtained Chem. Res. 45, 1014–1018.
DCPD-UPR-TO polymers by IR, 1 H NMR, and GPC indicates the Miyagawa, H., Mohanty, A.K., Burgueno, R., Drzal, L.T., Misra, M., 2007. Novel
biobased resins from blends of functionalized soybean oil and unsaturated
occurrence of the D–A reaction. Specially, GPC and SEM analysis
polyester resin. J. Polym. Sci. Polym. Phys. 45, 698–704.
demonstrate that TO molecules can completely react with DCPD- Pachpinyo, P., Lertprasertpong, P., Chuayjuljit, S., Sirisook, R., Pimpan, V., 2006.
UPR chains at a TO content up to 10%. By the incorporation of TO Preliminary study on preparation of unsaturated polyester resin/natural rub-
into DCPD-UPR, the thermo-mechanical properties, Tg , and stiffness ber latex blends in the presence of dispersion aids. J. Appl. Polym. Sci. 101,
4238–4241.
of the DCPD-UPR-TO polymer matrices decrease, while the ther- Pan, L.L., Li, G.Y., Su, Y.C., Lian, J.S., 2012. Fire retardant mechanism analysis between
mal stability (T10 and T50 ) and toughness of them are improved. ammonium polyphosphate and triphenyl phosphate in unsaturated polyester
The mechanical and thermal properties of the polymer matri- resin. Polym. Degrad. Stab. 97, 1801–1806.
Ray, D., 2009. Modification of the dynamic damping behavior of fly ash filled unsat-
ces are influenced by the synergistic effects of phase separation urated polyester resin/SBR latex blend composites. J. Reinf. Plast. Compos. 28,
and crosslink density. The mechanism analysis on properties will 1537–1552.

You might also like