Sei sulla pagina 1di 9

G Model

ARTICLE IN PRESS

INDCRO-7581; No. of Pages 9

Industrial Crops and Products xxx (2014) xxxxxx

Contents lists available at ScienceDirect

Industrial Crops and Products


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

Novel renewable resource-based UV-curable copolymers derived


from myrcene and tung oil: Preparation, characterization and
properties
Xuejuan Yang a , Shouhai Li a,b, , Jianling Xia a,b, , Jian Song a , Kun Huang a,b , Mei Li a
a
Institute of Chemical Industry of Forestry Products, CAF, Key Laboratory of Biomass Energy and Material, National Engineering Laboratory for Biomass
Chemical Utilization, Key and Laboratory on Forest Chemical Engineering, SFA, Nanjing 210042, Jiangsu Province, China
b
Institute of Forest New Technology, CAF, Beijing 100091, China

a r t i c l e

i n f o

Article history:
Received 25 July 2014
Received in revised form 13 October 2014
Accepted 14 October 2014
Available online xxx
Keywords:
Copolymers
Bio-based vinyl ester resins
Myrcene
Tung oil
UV-curable

a b s t r a c t
Two novel ultraviolet (UV)-curable vinyl ester resin (VER) monomers derived from myrcene and tung oil
were synthesized via DielsAlder reaction, glycidylation reaction and ring-opening esterication. Their
chemical structures were conrmed by Fourier transform-infrared analysis (FTIR) and Proton nuclear
magnetic resonance (1 H NMR). Then the two VER monomers were mixed at different weight ratios and
cured under UV light. The curing processes were monitored by FTIR analysis. Tensile and exural properties of the cured resins were also tested. The conversion rate of double bond exceeded 80% within 60 s and
the nal conversion rate was 95%. Tensile and exural tests indicated that the rigidityexibility was balanced in the cured systems when the mixed resins contained 5070 wt% of myrcene-based VER monomer.
Dynamic mechanical analysis demonstrated that the storage modulus and glass transition temperatures
of the cured resins increased with the increasing content of myrcene-based VER monomer. The crosslinking density of the cured copolymers also increased with the increasing content of myrcene-based VER
monomer. Thermogravimetric analysis (TGA) showed that the main thermal initial decomposition temperatures (Ti ) of the cured copolymers were all above 388 C. All the results demonstrated that comparable
balanced properties could be obtained by mixing these two bio-based monomers together.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Regarding the depletion of fossil fuels and the desire of environment protection, great efforts have been made in recent decades
to develop new polymers from renewable resources to replace
petroleum-based products. Agricultural and forestry feedstock are
the renewable resources of interest owing to their sufciency and
renewability (Shah et al., 2008; Chiari and Zecca, 2011). Now various bio-based polymers have been prepared, such as polyester,
polyurethanes, polyamides, epoxy resins and unsaturated resins
that are commonly used as thermosetting materials (Abdul Khalil
et al., 2010; Atta et al., 2007; Do et al., 2008).

Corresponding author at: Institute of Chemical Industry of Forestry Products,


CAF, Key Laboratory of Biomass Energy and Material, National Engineering Laboratory for Biomass Chemical Utilization, Key and Laboratory on Forest Chemical
Engineering, SFA, Nanjing 210042, Jiangsu Province, China. Tel.: +86 025 85482453;
fax: +86 025 85482454.
Corresponding author.
E-mail addresses: lishouhai1979@163.com (S. Li), xiajianling@126.com (J. Xia).

Vinyl ester resin, an important thermosetting resin, is usually a mixture of monomer, dimer and some other oligomers.
The curing mechanism of VER conforms to that of unsaturated
polyesters and its molecular structure similar to that of epoxy
resin (Robinette et al., 2004; Taillemite and Pauer, 2009). Therefore, cured VER exhibits excellent mechanical strength, chemical
resistance and perfect thermal stability, which make them widely
used in automotive industry, marine industry, metallurgy, military affairs, sports, pharmacy, construction and other industries
(Auad et al., 2001). Moreover, the UV-curable VER also owns high
rapid curing performance and therefore is used as surface coating, radiation curable inks and printed circuit board coatings (La
Scala et al., 2007; Sultania et al., 2010). In this work, we aim to
prepare two UV-curable VER monomers using myrcene and tung
oil as raw materials, respectively, and explore the comprehensive
properties of the copolymers with different weight ratios of these
two bio-based VERs.
Myrcene is a colorless or light yellow oily liquid that is mainly
extracted from the essential oils of many plants, such as Laurel, Verbena, Cypress and Hop. It contains two isomers, namely -myrcene
and -myrcene (Daferera et al., 2002), while -myrcene is rarely

http://dx.doi.org/10.1016/j.indcrop.2014.10.024
0926-6690/ 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model
INDCRO-7581; No. of Pages 9
2

ARTICLE IN PRESS
X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

studied owing to its instability in nature. Thus, the myrcene that


we usually referred to is -myrcene. In the industry, -myrcene is
mainly obtained from the pyrolysis of -pinene (Kolicheskia et al.,
2007). The chemical structure of myrcene contains three highly
reactive double bonds including a conjugated double bond. These
highly reactive double bonds make it can react with dienophile, H2 ,
O3 via DielsAlder reaction, hydrogenation or oxidation reaction
(Patricia et al., 2005; Goncalves et al., 2002; Atkinson et al., 1985).
One fragrance that could be used in deodorants with myrcene
as the starting material was recently synthesized, and AlCl3 was
the optimal catalyst for the addition of methanol to myrcene
(Behr et al., 2012). A high yielding myrac aldehyde issued from
myrcene was achieved via DielsAlder reaction between myrcene
and acrolein and with Zinc-containing ionic liquids as the catalysts.
As a result, the procedure is quite simple and convenient and the
products are very useful in perfumes and cosmetic powders (Yin
et al., 2005). Recently, we also synthesized a novel myrcene-based
VER monomer via DielsAlder reaction and ring-opening esterication with maleic anhydride and glycidyl methacrylate (Yang
et al., 2013). Despite high tensile strength and modulus, the UVcured product of the monomer is limited in exural strength and
impact property, which should be improved. As reported, myrcene
is traditionally used as synthetic fragrances and pharmaceutical
intermediates, but it has rarely been used as starting material in VER
elds, which hinders its further processing and utilization (Cawse
et al., 1987; Van den Berg et al., 1988). In this work, a bio-based
VER monomer was prepared via DielsAlder reaction, glycidylation reaction and epoxy ring-opening esterication using myrcene
as main raw material.
Tung oil is a triglyceride extracted from the seeds of tung tree
and consists of about 80% of polyunsaturated fatty acids chain,
named -eleostearic acids (i.e., 9-cis,11,13-trans-octad ecatrienoic
acid), and slightly of other acids (e.g. oleic acid and linoleic acid) (Li
et al., 2012). The large amount of conjugated triene system makes
it reactive with dienophile via DielsAlder reaction even without
any catalysts (Biermann et al., 2007). Various products have been
made from tung oil and widely used in coatings, varnishes, inks
and resins (Li and Larock, 2003; Oyman et al., 2005). An ester was
synthesized and used as the reactive diluents for alkyd resins via
transesterication of tung oil and alcohols. An excellent binding
agent system with lower viscosity and good curing properties was
obtained by using the ester as the diluents (Biermann et al., 2010).
Similarly, a tung-oil-based resin was synthesized and used as the
reactive toughening agent of the unsaturated polyester resin. The
toughness of the modied resin was improved obviously with the
increase of tung oil content, and the stiffnesstoughness was balanced when the content was 10% (Liu et al., 2013a). Two tung oil
based epoxies were also prepared and their performances were
superior to the epoxidized soybean oil, and especially the one has
more functional epoxy groups (Huang et al., 2013b). Some other
modied tung oils used in thermosetting resins were also investigated. For instance, tung oil reacted with bismaleimide at different
ratios to form a high-performance prepolymer and then after curing
at 200 C for 2 h, the cured products were homogeneous with excellent tensile strength of 38.1 MPa and modulus of 2.6 GPa (Shibata
et al., 2011). Tung oil also can be modied with maleic anhydride,
nonisocyanate polyurethane and acrylates to prepare UV-curable
resin (Huang et al., 2013). Tung oil is widely used in resins and performs extremely well in some areas (Liu et al., 2013b; Huang et al.,
2013a). In this work, we synthesized a UV-curable VER monomer by
polymerization of tung oil with acrylic acid via DielsAlder reaction,
followed by ring-opening esterication with glycidyl methacrylates (GMA).
Myrcene-based VER was a potential rigid material owing to the
alicyclic structure and high crosslinking density. However, tungoil-based VER behaved poorly with low strength, modulus and

glass transition temperature (Tg ) when used alone due to the long
aliphatic chain of triglyceride. Therefore, the two resins were mixed
at different weight ratios and cured under UV radiation using
dimethoxybenzoin as the photoinitiator in our experiments. We
also studied the curing behaviors, tensile properties, exural properties, dynamical mechanical properties and thermal stability of
the copolymers.
2. Experimental
2.1. Materials
Myrcene was purchased from Jiangxi Kaiyuan spices Co., Ltd,
contains about 75% of myrcene. Tung oil was obtained from the
Nanjing Daziran Fine chemicals Co. Ltd (stabilized, 95%, iodine
value is 167). Maleic anhydride (stabilized, 99.5%), hydroquinone
(stabilized, 99.5%) and Benzyltriethylamine Chloride (98.0%) were
purchased from The Group chemical reagent Co., Ltd. Epichlorohydrin (99.0%), acrylic acid (99.0%) and calcium oxide (99.5%)
were purchased from Shanghai Lingfeng chemical reagent Co., Ltd.
Sodium hydroxide (98.0%) and sodium bicarbonate (98.0%))
were purchased from Shanghai Zhanyun chemical Co., Ltd. Glycidyl methacrylate (98.5%) and petroleum ether (98.0%) were
purchased from Nanjing Kowloon Chemicals Co., Ltd. Dimethoxybenzoin (99%) was purchased from Aladdin Industrial Corporation.
All the reagents were used as received.
2.2. Synthesis
2.2.1. Synthesis of diglycidyl ester of maleinized myrcene
modied with acrylic acid (ADMM)
First, 150.00 g (1.53 mol) of maleic anhydride was charged to a
1000-mL ask equipped with a magnetic stirrer, a thermometer,
a dropping funnel and a reux condenser. The temperature was
raised to 55 C and kept until maleic anhydride melted completely.
Then 300.00 g (1.65 mol) of myrcene was added dropwise with the
dropping funnel while maintaining the temperature at 5560 C.
After the myrcene was added, the mixture was heated to 70 C and
reacted for 4 h. Next, the crude product was developed through
vacuum distillation (1.05 kPa) and the fraction of 190195 C was
collected (Goldblatt and Palkin, 1941). Finally, maleinized myrcene
was obtained as a light yellow liquid (yield: 91.10%).
Then 93.60 g (0.40 mol) of maleinized myrcene was added into
a 1000-mL ask equipped with a magnetic stirrer, a thermometer
and a reux condenser. With addition of 9.00 g (0.5 mol) of distilled water, the mixture reacted at 7080 C for 2 h until complete
hydrolysis. After that, 370.00 g (4.00 mol) of epichlorohydrin and
1.8 g (0.008 mol) of benzyltriethylamine chloride were added into
the ask. Then the mixture was heated to 100 C and the reaction continued at this temperature for 2 h. Then after the mixture
was cooled to 60 C, 32.00 g (0.80 mol) of sodium hydroxide and
44.80 g (0.80 mol) of calcium oxide were added under stirring at
this temperature for 3 h. At last, the product was ltered with celite
and the ltrate was collected. After the excessive epichlorohydrin
was distilled via rotary vacuum evaporation, a light yellowish viscous liquid, viz. diglycidyl ester of maleinized myrcene (DMM) was
obtained (yield: 87.90%).
Then 40.00 g (0.11 mol) of DMM, 0.28 g benzyltriethylamine
chloride and 0.17 g of hydroquinone were added to a ask equipped
with a magnetic stirrer, a thermometer, a dropping funnel and
a reux condenser. After the temperature was raised to 100 C,
15.85 g (0.22 mol) of acrylic acid was added slowly with the dropping funnel. The mixture reacted at 100 C for 3 h, and then a yellow
viscous resin (ADMM) was obtained (yield: 89.00%). The synthesis
route is shown in Fig. 1.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model
INDCRO-7581; No. of Pages 9

ARTICLE IN PRESS
X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

Fig. 1. The synthesis route of ADMM.

2.2.2. Synthesis of tung oil acrylic acid (TOA) polymerized


glycidyl methacrylate (TOA-GMA)
First, 44.10 g (0.05 mol) of tung oil and 0.0856 g of hydroquinone
were added to a ask equipped with a magnetic stirrer, a thermometer, a dropping funnel and a reux condenser. After heating
to 180 C, 12.96 g (0.18 mol) of acrylic acid was added dropwise to
reacted for 5 h. Then the excessive acrylic acid was removed by
vacuum distillation (1.05 kPa) and a transparent light orange liquid of TOA was obtained (acid value is 150 mg/g, theory value is
153 mg/g).
Next, 33.00 g (0.0295 mol) of acrylic acid modied tung oil,
0.114 g hydroquinone and 0.228 g hexadecyl trimethyl ammonium
chloride were added into a ask equipped with a magnetic stirrer,
a thermometer, a dropping funnel, an inert gas inlet and a reux
condenser. After the temperature was raised to 120 C, 12.57 g
(0.0884 mol) of glycidyl methacrylate was added slowly with the
dropping funnel and the reaction continued for about 1 h until the
acid value dropped below 3 mg/g. TOA-GMA was obtained as an
orange viscous liquid. The synthesis route is shown in Fig. 2.
2.3. Preparation of cured samples
ADMM and TOA-GMA were mixed uniformly by different
weight ratios of 0%/100%, 30%/70%, 50%/50%, 70%/30% and 100%/0%,
respectively. Then 3 wt% (on the basis of total weight of resins)
of dimethoxybenzoin was added as the photoinitiator. Cured
polymer samples were prepared by casting the above mixture into molds with dimensions of 80 mm 10 mm 4 mm and
7.62 mm 3.18 mm 0.5 mm. Then the curing was performed in
the Intelli Ray 400UV light curing box for 20 min with the irradiation intensity of 100 mW/cm2 . The copolymers were labeled
as follows: ADMM30/TOA-GMA70 means a mixture of 30 wt% of
ADMM and 70 wt% of TOA-GMA.

of the instrument is better than 0.4 cm1 , the accuracy of wave


number is 0.01 cm1 . Samples were analyzed as liquid lms on a
ZnSn window. Each of samples was scanned from 4000 to 400 cm1 .
1 H NMR (300 MHz) spectra were recorded on an ARX300 spectrometer (Bruker, Germany). The chemical shifts relative to that of
deuterated chloroform (d 7.26) were recorded.
The acid value of resin was measured by titration of 0.5 g resin
in acetone solution by 0.1 N sodium hydroxide solution and phenolphthalein was used as the indicator (GB/T 2895-2008).
Flexural property was measured using a CMT4303 universal testing machine (SANS, China) equipped 1 kN electronic load
cell according to ASTM D790. And the tests were conducted at
a crosshead speed of 1 mm/min with a support span of 50 mm.
Tensile properties were also measured using CMT4303 universal
testing machine (SANS, China) following ASTM D638-03. For the
sake of accuracy, ve replicates were measured and the average
values were obtained. All the samples were tested at 25 C.
The dynamic mechanical analysis (DMA) was performed on a
Q800 dynamic mechanical thermal analyzer (TA, USA) in threepoint bending mode with a frequency of 1 Hz. All samples had a
dimension of 60 mm 10 mm 4 mm and were swept from 40
to 120 C at a heating rate of 3 C/min.
Thermogravimetric analysis (TGA) was performed using a 409PC
thermogravimetric analyzer (Netzsch, Germany). Each of samples
was tested from 25 to 800 C at a heating rate of 15 C/min under
a nitrogen atmosphere. After a detail calculation and analysis, the
values of Ti and Tmax can be obtained. Ti is the corresponding temperature of intersection of the extension cord of platform at low
temperature and the tangent through the inection point of main
thermal degradation region on TG curve. Tmax is the corresponding
temperature of inection point of main thermal degradation region
on TG curve.
3. Results and discussion

2.4. UV curing behaviors of the mixed monomers


The mixtures prepared as above were evenly smeared on the
clean glass plates with a 50 m obliterator and cured under UV light
(100 mW/cm2 ). Each of cured samples was collected at 0, 10, 30, 60,
120, 240, 480 s, respectively. The cured behaviors were investigated
by FTIR analysis. The absorption peak intensity of the double bonds
at about 1635 cm1 and carbonyl group at about 1725 cm1 were
calculated, respectively. And the conversion rate of double bond
can be obtained by contrasting the integral area of the two peaks.
2.5. Characterization
FTIR analysis was performed using an IS10 spectrometer (Nicolet, USA) by an attenuated total reectance method. The resolution

3.1. Synthesis and characterization of ADMM and TOA-GMA resin


monomers
3.1.1. FTIR analysis of ADMM and TOA-GMA monomers and their
intermediates
The FTIR spectra of myrcene based VER monomer and its intermediates are shown in Fig. 3. The spectrum of myrcene showed
several characteristic peaks: the conjugated double bonds (1644
and 1594 cm1 ), terminal vinyl group (3089 and 1796 cm1 ), and
isolated double bond (1644 cm1 ). All these absorption peaks
clearly veried the molecular structure of myrcene. The spectrum
of maleinized myrcene also showed some typical peaks: carbonyl of
the cyclic anhydride (1834 and 1768 cm1 ), and bending vibration
of C H in the hexatomic ring (1311 cm1 ). The absorption peaks of

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model
INDCRO-7581; No. of Pages 9
4

ARTICLE IN PRESS
X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

Fig. 2. The synthesis route of TOA-GMA.

conjugated double bonds and terminal vinyl group all disappeared.


These changes indicated that maleinized myrcene was already synthesized successfully. On the FTIR spectrum of DMM, the absorption
peaks of cyclic anhydride at 1834 and 1768 cm1 disappeared and
a new intensity absorption peak of carbonyl group at 1732 cm1
appeared after the reaction between maleinized myrcene and epichorohydrin. And the characteristic peaks of C O C (1439 and
1176 cm1 ) and epoxy group (908 cm1 ) all appeared. On the spectrum of ADMM, the absorption peak of epoxy group disappeared.
In addition, the peaks of free hydroxyl group (3460 cm1 ), double

bond (1635 cm1 ) and the multiple absorption peaks of ester carbonyl conjugated with vinyl group all appeared. These changes and
the appearance of the groups demonstrated that the acrylic acid
already reacted with DMM via ring-opening esterication.
The FTIR spectra of tung oil-based VER monomer and its intermediates are showed in Fig. 4. The spectrum of tung oil displayed
several characteristic absorption peaks: the conjugated triene
(3014, 987 and 964 cm1 ), methyl and methylene groups (2924
and 2853 cm1 ) and ester carbonyl groups (1738 cm1 ). On the FTIR
spectrum of TOA, the absorption peaks of the conjugated triene

Fig. 3. The FTIR spectra of myrcene-derived VER monomer and its intermediates.

Fig. 4. The FTIR spectra of tung oil-based VER monomer and its intermediates.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model
INDCRO-7581; No. of Pages 9

ARTICLE IN PRESS
X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

Fig. 6. UV-curing behaviors of mixed VER monomers with different weigh ratios.

group. The appearance of the peak at about 3.70 ppm further indicated the occurrence of ring-opening esterication. The peaks at
2.90 and 2.33 ppm corresponded to the protons in the hexatomic
ring, which conrmed the occurrence of the DielsAlder reaction.
All the evidences in the 1 H NMR spectrum of ADMM veried the
chemical structure of the ADMM monomer.
For the 1 H NMR spectrum of TOA-GMA, the corresponding characteristic peaks of all the protons were also displayed clearly.
The characteristic chemical shifts at 2.80, 2.70 and 2.10 ppm corresponded to the protons on the hexatomic ring formed via
DielsAlder reaction between tung oil and acrylic acid. The chemical shifts at 3.77 and 3.50 ppm were attributed to the protons
on the CH2 connected with hydroxyl, which was generated
by the esterication of glycidyl methacrylate. The appearance
of all chemical shifts indicated that TOA-GMA was successfully
synthesized.
Fig. 5.

HNMR spectra of ADMM and TOA-GMA monomers.

disappeared.
The
typical
wide
absorption
bands
at
24503450 cm1 appeared. The existence of the hydrogen
bond association made the absorption peaks wide and overlapped
with the peaks of methyl and methylene. Several small peaks
appeared at about 1310 cm1 , which corresponded to the bending
vibration of C H in the CH2 of the hexatomic ring. The peaks
at 1740 and 1703 cm1 was assigned to the C O of ester and
carboxyl group, respectively. On the FTIR spectrum of TOA-GMA,
the absorption peaks of C O (1703 cm1 ) and associated hydroxyl
(24503450 cm1 ) that ever appeared in the FTIR spectrum of
TOA disappeared. And the free hydroxyl group (3483 cm1 ) was
produced by the ring-opening esterication. The peak intensity of
C O at 1720 cm1 increased greatly compared with that of tung
oil. All the above evidences demonstrated that the target product
was successfully synthesized.
3.1.2. 1 H NMR analysis of ADMM and TOA-GMA VER monomers
The 1 H NMR spectra of ADMM and TOA-GMA are shown in Fig. 5.
For ADMM, the chemical shift from 5.70 to 6.53 ppm was attributed
to the protons of acrylic acid part, and the chemical shifts from
4.89 to 5.46 ppm were assigned to the protons of CH CH2 in the
myrcene part. In addition, the chemical shift from 3.40 to 4.46 ppm
corresponded to the protons of glycidy ester, which were produced
after ring-opening esterication between acrylic acid and glycidyl

3.2. UV-curing behaviors of ADMM and TOA-GMA mixed


monomers
The UV-curing behaviors of copolymers with different weight
ratios are displayed in Fig. 6. Clearly, the nal double bond conversion rates of different copolymers were all above 84% after
480 s of curing, indicating that the copolymerized reactivity of
the two monomers was excellent. The conversion rates of double bond of all copolymers increased rapidly and were up to 80%
within 60 s (Fig. 6). However, after 60 s, the conversion rate proceeded slowly as the viscosity was enhanced with the elevated
curing degree. The nal conversion rates of double bond of cured
copolymers were improved with the increasing ADMM content
and they were all higher than that of the homopolymers of the
two resins. This improvement was attributed to the higher reactivity and lower viscosity of ADMM that made the monomer radical
more movable. The viscosity of TOA-GMA was higher than that
of ADMM as it contained more branched-chains, and the more
severe intermolecular interactions hindered the resin monomer
from free movement. On the other hand, for pure ADMM VER
monomer, the big and rigid alicyclic structure also prevented
the molecular chain from free movement, which also hindered
the copolymerization of double bonds. These two factors jointly
inuenced the nal double bond conversion rates of cured copolymers.
When ADMM contents were 50 wt% and 70 wt%, their curing
curves intersected at certain curing time points. At the beginning,

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model
INDCRO-7581; No. of Pages 9
6

ARTICLE IN PRESS
X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

3.3. Tensile and exural properties of the cured polymers

Fig. 7. Tensile stressstrain curves of copolymers with different weight ratios.

the curing rate of the copolymer with 50 wt% of ADMM was slightly
higher while its nal double bond conversion rate was lower than
that of the copolymer with 70 wt% of ADMM. This was probably
because at the beginning, the viscosity did not greatly affect the
curing process. As the curing went on, the molecular chains crosslinked, the double bond density decreased sharply and the curing
system was more diffusion-controlled. Thus, the copolymer with
higher ADMM content resulted in a higher nal double bond conversion rate. We can see from the trends of curing curves that when
ADMM content was between 50% and 70%, the curing rate and double bond conversion rate of copolymer were relatively higher. The
nal double bond conversion rate was up to 95%.

Fig. 7 shows the tensile stressstrain curves of the cured resins


with different ADMM/TOA-GMA weight ratios. The detailed data
of tensile and exural properties are summarized in Table 1.
Clearly, the breaking elongation of copolymers ranged from 3.2%
to 14.9% and tensile strength ranged from 49.0 to 4.8 MPa. The
cured resin with 100% ADMM exhibited the highest tensile strength
and the lowest breaking elongation, while the cured pure TOAGMA showed the highest breaking elongation and lowest tensile
strength. These results could be explained by the fact that compared
with TOA-GMA, the density of the rigid alicyclic hexatomic and
the crosslinking density of double bonds in the ADMM resin were
higher, which bestowed the cured system with higher strength.
In addition, the aliphatic chains that could enhance the exibility
of the cured resin were shorter in ADMM than in TOA-GMA. All
these factors endowed the cured pure ADMM resin with excellent
tensile strength. On the contrary, TOA-GMA monomer contained
more exible aliphatic chains and greater amount of free volume.
Thus, when the cured sample was stretched, the long chain moved
more freely to offset the external forces and reach a balance, which
bestowed the cured TOA-GMA with higher ability to extend, and
thus resulted in a decrease of modulus.
The tensile strength of copolymers with different ADMM/TOAGMA weight ratios was enhanced with the increasing ADMM
content, while the breaking elongation rst increased with the
increasing ADMM content to 60% and then decreased (Fig. 7). The
tensile properties of the cured resins are determined by the chemical structures and the cross-linked states of the resins. The chemical
structures and cross-linked states in the cured copolymers varied
along with the ratios of monomers. When the TOA-GMA content
was high, a small part of TOA-GMA copolymerized with ADMM
monomer and the most part was homopolymerized. The ADMM
monomer acted as the point of junction, and thus the copolymer exhibited the similar properties to the pure cured TOA-GMA

Fig. 8. The possible crosslinking schematic diagram of the cured resins.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model

ARTICLE IN PRESS

INDCRO-7581; No. of Pages 9

X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

Table 1
Tensile and exural properties of the cured copolymers with different weight ratios.
Formula

Tensile properties
Strength (MPa)

ADMM
ADMM70/TOA-GMA30
ADMM60/TOA-GMA40
ADMM50/TOA-GMA50
ADMM40/TOA-GMA60
ADMM30/TOA-GMA70
TOA-GMA

49.0
33.9
28.4
23.7
20.4
15.6
4.8

2.1
1.5
0.8
0.6
0.9
0.8
0.5

Flexural properties
Breaking elongation (%)
3.2
8.1
10.9
9.5
8.3
7.7
14.9

resin. When ADMM content increased, the density of alicyclic hexatomic structure also increased, which contributed to improve the
strength and modulus of copolymers. When ADMM content was up
to 5070 wt%, it played a dominate role in the copolymer. At this
weight ratio, ADMM acted as the main resin, and TOA-GMA was the
ductile connection point. Under this circumstance, the mixed resins
formed hyper-branched and star-shaped copolymer (Fig. 8, KargerKocsis et al., 2004). The copolymer combined the high strength,
modulus and breaking elongation of these two resins. Although
the exural strength was slightly lower than that of the acrylated
epoxidized soybean oil (AESO) and vinyl ester copolymer (61 MPa)
reported by Grishchuk, in whose work the weight ratio of vinyl
ester and AESO was 50/50, the modulus was much higher than the
vinyl ester and AESO copolymer (1516 MPa) (Grishchuk and KargerKocsis, 2011). The breaking elongation of the ADMM/TOA-GMA
copolymer was also much higher than the traditional petroleumbased vinyl ester resin, whose breaking elongation was generally
lower than 5%.

0.2
0.3
0.5
0.3
0.3
0.2
0.5

Strength (MPa)
98.8
61.9
56.1
46.0
34.1
28.9
5.5

3.5
2.1
2.0
1.9
1.8
1.8
0.3 (unbreak)

Modulus (MPa)
3657.6
1795.1
1942.1
1642.2
853.9
842.0
154.3

50.1
30.6
31.9
25.6
15.3
16.1
6.9

Besides, there were many possible relaxation modes in the longer


aliphatic chain of TOA-GMA, which resulted in the slight broadening of Tg peak (La Scala and Wool, 2005). Based on the above results
we can infer that the TOA-GMA works as the reactive plasticizer in
the copolymer system.
E is also closely related to the crosslinking density, which can
be determined according to the rubber elasticity theory. Thus the
experimental crosslinking density (e ) of the cured resins can be
calculated as follows (Asif et al., 2005; Liu et al., 2013a; Lu et al.,
2005):
E  = 3e RT
where E is the storage modulus of the cross-linked polymer in the
rubber state; R is the gas constant and T is the absolute temperature.

3.4. Dynamic mechanical analysis (DMA) of copolymers


Fig. 9 shows the storage modulus (E ) and loss factor (tan)
curves of cured resins with different ADMM/TOA-GMA weight
ratios. The peak temperature of tan corresponded to the glass transition temperature (Tg ). Clearly, the E of the copolymers increased
with the increasing ADMM content. E was highest in the cured
pure ADMM, but was the lowest in the pure TOA-GMA owing to its
less rigid alicyclic structure, lower cross-linked density and longer
aliphatic chains. The exible aliphatic chains were replaced by the
rigid alicyclic structure with the increasing ADMM content, which
reduced the free volume in the polymeric matrix. On the other hand,
ADMM monomer was more reactive than TOA-GMA monomer as
the unsaturated sites in the middle of the fatty acid chains would
reduce the curing rate. These factors accounted for the changing
trend of E mentioned above.
The tan curves showed that each of the cured resin had one
clearly single Tg , which demonstrated that all the cured systems
were compatible and homogenous. The Tg of the pure cured ADMM
resin was 77.1 C, then it decreased with the decreasing ADMM
content, and the Tg of the cured pure TOA-GMA was 34.8 C. As
reported, the Tg of ADMM/TOA-GMA copolymer prepared in our
study was signicantly higher than that of the biobased copolymer reported by Campanellas research team. In their study, 33 wt%
of methacrylated lauric acid was used as a comonomer to copolymerized with AESO, while the Tg of the methacrylated lauric acid
andAESO copolymer was only 27 C without blending with styrene
(Campanella et al., 2011). Thus, ADMM/TOA-GMA was a potential material in the VER area. As is well-known, Tg of a cured
resin is also affected by the structure of the monomers and the
cross-linked state as previously discussed. Moreover, the tan peak
became less intensive and slightly broader with the decreasing
ADMM content. The intensity reduction of the Tg peak was due to
the hyper-branched and star-shaped structure of the copolymer.

Fig. 9. DMA curves of copolymers with different weight ratios.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model

ARTICLE IN PRESS

INDCRO-7581; No. of Pages 9

X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

Table 2
The crosslinking density of copolymers with different weight ratios.

Table 3
TGA data of cured resins with different ADMM/TOA-GMA weight ratios.

Formula

Tg ( C)

E at Tg + 20 C (MPa)

e (mol/m3 )

Formula

T5% ( C)

Ti ( C)

Tmax ( C)

Char yield (%)

ADMM
ADMM70/TOA-GMA30
ADMM50/TOA-GMA50
ADMM30/TOA-GMA70
TOA-GMA

77.1
65.1
58.0
52.5
34.8

35.7
37.5
38.9
39.3
31.0

3866
4197
4441
4559
3790

ADMM
ADMM70/TOA-GMA30
ADMM50/TOA-GMA50
ADMM30/TOA-GMA70
TOA-GMA

357.2
289.5
282.4
282.4
266.1

402.9
388.7
388.9
388.0
382.9

439.8
426.4
430.8
433.1
422.5

0.3
4.2
3.1
2.6
5.5

In this study, the E at Tg + 20 C was used, which ensured that the


cured resins were in the rubber state (Can et al., 2006). The data of
the crosslinking density of the cured systems are shown in Table 2.
Clearly, e of the copolymers was higher than that of the pure
ADMM and TOA-GMA resins. Theoretically, TOA-GMA and ADMM
have three and two reactive terminal double bonds, respectively.
TOA-GMA has a molecular weight 3.5 times higher than ADMM.
Thus, the total number of cross-linkable points in copolymerized
system increased with the increasing ADMM content (Campanella
et al., 2009). However, when copolymerized system has a high
content of ADMM, excessive ADMM molecules could form a certain length of cross-linking bridge covalently joining all molecular
chains together, which in turn results a reduction of cross-linked
density (Li et al., 2013). Thus, an increase in ADMM content resulted
in lower cross-link density.

3.5. Thermogravimetric analysis (TGA) of copolymers


The thermal stability analysis results of cured resins with different ADMM/TOA-GMA weight ratios are shown in Fig. 10. The char
yield at the end of the TGA, and the temperature at which the weight
loss occurred are also listed in Table 3. Clearly, degradation of all the
cured products was a two-steps thermal process with a small degradation and a large one. And the Ti of the main degradation region
of all the cured resins exceeded 380 C. The Ti and Tmax of the two
mixed resins with different ADMM/TOA-GMA weight ratios were
very close, but were higher compared with pure TOA-GMA and
lower compared with pure ADMM. This was because there existed
numerous conned rigid alicyclic molecular structure in pure cured
ADMM, which can retards the decomposition and subsequent diffusion of decomposition products, thus it endowed the pure cured
ADMM with high thermostability. When the ADMM/TOA-GMA
weight ratio was 70/30, the T5% (temperature corresponding to 5%
of weigh loss) of cured resin was 289.5 C, which was comparably
higher than other mixed resins. The char yield at the end of the test
was 4.2%, which was also higher compared with other mixed resins.
At this weight ratio, a good cross-linked system was formed. This
result also agreed with the mechanical properties tests and DMA
results.

4. Conclusions

Fig. 10. TGA curves of copolymers with different weight ratios.

Two novel bio-based VER monomers, ADMM and TOA-GMA,


were prepared. The pure cured ADMM was proved to be a rigid
material with high tensile strength, modulus and thermal stability, but the cured pure tung oil-based VER had lowest strength
and modulus. Thus, we copolymerized two VERs with different
weight ratios to improve the toughness of myrcene-derived VER
and enhance the strength of the tung oil-derived VER.
The UV curing behavior analysis indicated that their nal double conversion rates were up to 90%. Tensile, exural tests and
DMA results revealed that the tensile strength, exural strength,
Tg and the storage modulus (E ) were improved with the increasing
ADMM content, while the breaking elongation and exural modulus rst increased and then decreased. TGA results showed that
the main thermal initial degradable temperatures of the copolymer
were all above 388 C and very constant. The char yield decreased
and the temperature corresponding to 5% weight loss increased
with the increasing ADMM content. The results indicated that when
the mixed resin contained 5070 wt% of ADMM, the cured resins
reached a comparable balanced performance.
This study presents a complementation between ADMM and
TOA-GMA and demonstrates that myrcene and tung oil were
successfully used to prepare fully bio-based novel materials
with desired properties. Besides, the copolymer used renewable
resources as raw materials instead of blending with styrene, and
thus, it has the advantage of no volatile poison emissions. Therefore,
the novel fully bio-based VER with superior mechanical properties
and promising thermal stability can partially replace the conventional VER-based thermosetting plastics.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

G Model
INDCRO-7581; No. of Pages 9

ARTICLE IN PRESS
X. Yang et al. / Industrial Crops and Products xxx (2014) xxxxxx

Acknowledgements
The authors are grateful to the nancial support from
Jiangsu Province Science Foundation for Youths (grant number:
BK2012065) and Chinese Nature Science Foundation (grant number: 31170544).
References
Abdul Khalil, H.P.S., Nurul Fazita, M.R., Bhat, A.H., Jawaid, M., Nik Fuad, N.A., 2010.
Development and material properties of new hybrid plywood from oil palm
biomass. Mater. Des. 31, 417424.
Asif, A., Shi, W., Shen, X., Nie, K., 2005. Physical and thermal properties
of UV curable waterborne polyurethane dispersions incorporating hyperbranched aliphatic polyester of varying generation number. Polymer 46,
1106611078.
Atkinson, R., Aschmann, S.M., Winer, A.M., Pitts, J.N., 1985. Kinetics and atmospheric implications of the gas-phase reactions of nitrate radicals with a series
of monoterpenes and related organics at 294 2 K. Environ. Sci. Technol. 19,
159163.
Atta, A.M., Nassar, I.F., Bedawy, H.M., 2007. Unsaturated polyester resins based on
rosin maleic anhydride adduct as corrosion protections of steel. React. Funct.
Polym. 67, 617626.
Auad, M.L., Frontini, P.M., Borrajo, J., Aranguren, M.I., 2001. Liquid rubber
modied vinyl ester resin, fracture and mechanical behavior. Polymer 42,
37233730.
Behr, A., Johnen, L., Neubert, P., 2012. A sustainable route from the renewable
myrcene to methyl ethers via direct hydroalkoxylation. Catal. Sci. Technol. 2,
8892.
Biermann, U., Butte, W., Holtgrefe, R., Feder, W., Metzger, J.O., 2010. Esters of calendula oil and tung oil as reactive diluents for alkyd resins. Eur. J. Lipid Sci. Technol.
112, 103109.
Biermann, U., Butte, W., Eren, T., Haase, D., Metzger, J.O., 2007. Regio- and stereoselective DielsAlder additions of maleic anhydride to conjugated triene fatty acid
methyl esters. Eur. J. Org. Chem. 23, 38593862.
Campanella, A., Scala, J.J.L., Wool, R.P., 2011. Fatty acid-based comonomers as styrene
replacements in soybean and castor oil-based thermosetting polymers. J. Appl.
Polym. Sci. 119, 10001010.
Campanella, A., La Scala, J.J., Wool, R.P., 2009. The use of acrylated fatty acid methyl
esters as styrene replacements in triglyceride-based thermosetting polymers.
Polym. Eng. Sci. 49, 23842392.
Can, E., Wool, R.P., Ksefoglu, S., 2006. Soybean- and castor-oil-based thermosetting
polymers: mechanical properties. J. Appl. Polym. Sci. 102, 14971504.
Cawse, J.L., Stanford, J.L., Still, R.H., 1987. Polymers from renewable sources: 5.
Myrcene-based polyols as rubber-toughening agents in glassy polyurethanes.
Polymer 28, 368374.
Chiari, L., Zecca, A., 2011. Constraints of fossil fuels depletion on global warming
projections. Energy Policy 39, 50265034.
Daferera, D., Pappas, C., Tarantilis, P.A., Polissiou, M., 2002. Quantitative analysis of
-pinene and -myrcene in mastic gum oil using FT-Raman spectroscopy. Food
Chem. 77, 511515.
Do, H.S., Park, J.H., Kim, H.J., 2008. UV-curing behavior and adhesion performance of
polymeric photoinitiators blended with hydrogenated rosin epoxy methacrylate for UV-crosslinkable acrylic pressure sensitive adhesives. Eur. Polym. J. 44,
38713882.
Goldblatt, L.A., Palkin, S., 1941. Vapor phase thermal isomerization of -and pinene. J. Am. Chem. Soc. 63, 35173522.
Goncalves, J.A., Howarth, O.W., Gusevskaya, E.V., 2002. Palladium catalyzed oxidation of monoterpenes: novel oxidation of myrcene with dioxygen. J. Mol. Catal.
A: Chem. 185, 97104.
Grishchuk, S., Karger-Kocsis, J., 2011. Hybrid thermosets from vinyl ester resin and
acrylated epoxidized soybean oil (AESO). Express Polym. Lett. 5, 211.

Huang, K., Zhang, J., Li, M., Xia, J., Zhou, Y., 2013a. Exploration of the complementary
properties of biobased epoxies derived from rosin diacid and dimer fatty acid
for balanced performance. Ind. Crop Prod. 49, 497506.
Huang, K., Zhang, P., Zhang, J., Li, S., Li, M., Xia, J., Zhou, Y., 2013b. Preparation of
biobased epoxies using tung oil fatty acid-derived C21 diacid and C22 triacid
and study of epoxy properties. Green Chem. 15, 24662475.
Huang, Y., Pang, L., Wang, H., Zhong, R., Zeng, Z., Yang, J., 2013. Synthesis and properties of UV-curable tung oil based resins via modication of DielsAlder reaction,
nonisocyanate polyurethane and acrylates. Prog. Org. Coat. 76, 654661.
Karger-Kocsis, J., Frhlich, J., Gryshchuk, O., Kautz, H., Frey, H., Mlhaupt, R., 2004.
Synthesis of reactive hyperbranched and star-like polyethers and their use for
toughening of vinylesterurethane hybrid resins. Polymer 45, 11851195.
Kolicheskia, M.B., Coccob, L.C., Mitchellc, D.A., Kaminski, M., 2007. Synthesis of
myrcene by prolysis of -pinene: analysis of decomposition reactions. J. Anal.
Appl. Pyrolysis 80, 92100.
La Scala, J., Wool, R.P., 2005. Property analysis of triglyceride-based thermosets.
Polymer 46, 6169.
La Scala, J.J., Ulven, C.A., Orlicki, J.A., Jain, R., Palmese, G.R., Vaidya, U.K., Sands, J.M.,
2007. Emission modeling of styrene from vinyl ester resin. Clean Technol. Environ. Policy 9, 265279.
Li, F., Larock, R.C., 2003. Synthesis, structure and properties of new tung
oil-styrene-divinylbenzene copolymers prepared by thermal polymerization.
Biomacromolecules 4, 10181025.
Li, M., Xia, J.L., Li, S.H., Huang, K., Wang, M., 2012. Study on fatty acid composition
and variation analysis of tung oils in China by GC/MS. Adv. Mater. Res. 554,
20182023.
Li, S., Xia, J., Li, M., Huang, K., 2013. New vinyl ester bio-copolymers derived from
dimer fatty acids: preparation, characterization and properties. J. Am. Oil. Chem.
Soc. 90, 695706.
Liu, C., Lei, W., Cai, Z., Chen, J., Hu, L., Dai, Y., Zhou, Y., 2013a. Use of tung oil as a reactive toughening agent in dicyclopentadiene-terminated unsaturated polyester
resins. Ind. Crop Prod. 49, 412418.
Liu, C., Dai, Y., Wang, C., Xie, H., Zhou, Y., Lin, X., Zhang, L., 2013b. Phase-separation
dominating mechanical properties of a novel tung-oil-based thermosetting
polymer. Ind. Crop Prod. 43, 677683.
Lu, J., Khot, S., Wool, R.P., 2005. New sheet molding compound resins from soybean
oil. I. Synthesis and characterization. Polymer 46, 7180.
Oyman, Z.O., Ming, W., Linde, R.V.D., 2005. Oxidation of drying oils containing nonconjugated and conjugated double bonds catalyzed by a cobalt catalyst. Prog.
Org. Coat. 54, 198204.
Patricia, A.R.D., Marcelo, G.S., Edsia, M.B.S., Eduardo, N.D.S., Elena, V.G., 2005. Slective hydrogenation of myrcene catalyzed by solgel Pd/SiO2 . Appl. Catal. A: Gen.
295, 5258.
Robinette, E.J., Ziaeel, S., Palmese, G.R., 2004. Toughening of vinyl ester resin using
butadiene-acrylonitrile rubber moditiers. Polymer 45, 61436154.
Shah, A.A., Hasan, F., Hameed, A., Ahmed, S., 2008. Biological degradation of plastics:
a comprehensive review. Biotechnol. Adv. 26, 246265.
Shibata, M., Teramoto, N., Nakamura, Y., 2011. High performance bio-based thermosetting resins composed of tung oil and bismaleimide. J. Appl. Polym. Sci.
119, 896901.
Sultania, M., Rai, J.S.P., Srivastava, D., 2010. Studies on the synthesis and curing of
epoxidized novolac vinyl ester resin from renewable resource material. Eur.
Polym. J. 46, 20192032.
Taillemite, S., Pauer, R., 2009. Bright future for vinyl ester resin in corrosion applications. Reinf. Plast. 53, 3437.
Van den Berg, K.J., Van den Horst, J., Boon, J., Sudeiijer, J.O.O., 1988. Cis-1,4-poly-myrcene: the structure of the polymeric fraction of mastic resin (Pistacia lentiscus
L.) elucidated. Tetrahedron Lett. 39, 26452648.
Yang, X.J., Li, S.H., Tang, X.D., Xia, J.L., 2013. Research on preparation, properties and
UV curing behaviors of novel myrcene-based vinyl ester resin. Adv. Mater. Res.
807, 28262830.
Yin, D., Li, C., Li, B., Tao, L., Yin, D., 2005. High regioselective DielsAlder reaction of
myrcene with acrolein catalyzed by zinc-containing ionic liquids. Adv. Synth.
Catal. 347, 137142.

Please cite this article in press as: Yang, X., et al., Novel renewable resource-based UV-curable copolymers derived from myrcene and
tung oil: Preparation, characterization and properties. Ind. Crops Prod. (2014), http://dx.doi.org/10.1016/j.indcrop.2014.10.024

Potrebbero piacerti anche