Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Industrial Crops and Products 36 (2012) 485–499

Contents lists available at SciVerse ScienceDirect

Industrial Crops and Products


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

The development and comparison of bio-thermoset plastics from epoxidized


plant oils
Joo Ran Kim, Suraj Sharma ∗
Department of Textiles, Merchandising and Interiors, University of Georgia, 350 Dawson Hall, Athens, GA 30602, USA

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

Article history: This study investigates the development of thermoset plastics from plant-based oils (e.g., linseed, soy-
Received 4 May 2011 bean, cottonseed, oilseed radish, and peanut oils) using an optimal process of solvent-free epoxidation.
Received in revised form 24 October 2011 The epoxidation of plant oils can be accomplished economically by reacting the double bonds of fatty-
Accepted 26 October 2011
acids with hydrogen peroxide. During the solvent-free process catalyzed by the ion-exchange resin, we
Available online 3 December 2011
observed that the influence of several variables was important: the molar ratio of hydrogen peroxide to
unsaturation, acetic acid to unsaturation, and temperature. The epoxidation of plant oils was determined
Keywords:
from the liquid mixture and the composite matrix by thermal and spectroscopic analyses. Compounds
Epoxidation
Linseed
with a higher double-bond (iodine) value showed higher oxirane oxygen percent and selectivity, and
Soybean a higher hydroxyl value because of a greater possibility of attack by solutions causing side reactions.
Oilseed radish Lower iodine values indicated fewer epoxy groups and selectivity, and a lower hydroxyl value. Benzyl
Cottonseed pyrazinium hexafluoroantimonate (BPH) yielded good thermal curing properties; as little as 1% added
Peanut to the plastics produced light-weight composites. Epoxidized linseed oil promises the highest modulus
DMA and impact resistance due to the largest number of double bonds to contribute more epoxy groups and
BPH the large proportion of linolenic acids to produce epoxy groups rapidly.
FA composition
© 2011 Elsevier B.V. All rights reserved.

1. Introduction or H2 O2 (hydrogen peroxide), and are considered an inexpensive


renewable material for the generation of biobased thermoset epoxy
Thermoset composites from epoxy resins are often used in resins (Boquillon and Fringant, 2000). Many plant oils process have
high-performance structural applications because they generally been studied to develop epoxy functionalized resins into the bio-
possess excellent properties such as toughness and dimensional based thermoset polymers: soybeans (Zaher et al., 1989), linseed
stability (Adachi et al., 2002). These resins are generally used as the (Chen et al., 2002), canola (Mungroo et al., 2008), and cottonseed
binding agent, offering attractive combinations of strength, ease (Dinda et al., 2008).
of processing, and cost. For example, epoxy resin is used as the In general, four methods are available for epoxidation of oils:
binder in composites reinforced with carbon fibers for aerospace (i) epoxidation with percarboxylic acids by acids or enzymes; (ii)
applications (Williamson, 2001). However, in general these resin epoxidation with inorganic and organic peroxides which includes
are derived from petroleum based chemicals. Therefore, several alkaline and nitrile hydrogen peroxide; (iii) epoxidation with halo-
research studies are being undertaken to explore the alternative, hydrins; and (iv) epoxidation with molecular oxygen (Guenter
sustainable biopolymers for their use in thermoset applications. et al., 2003). Epoxidation of oils occurs in three stages: (i) for-
Bio-based materials are finding increased attention due to their mation of peroxyacetic acid, (ii) reaction of the peroxyacetic acid
availability on a renewable basis and associated environmental and the double bonds, and (iii) side reactions which hydrolyze the
benefits. Plant oils, which are renewable resources, are found in epoxy group, resulting in degradation of the epoxy group as shown
abundance and have the potential to replace the petroleum-based schematically in Fig. 1(A and B) (Rangarajan et al., 1995; Rüsch gen
polymers (Can et al., 2006; Zhao et al., 2008). Natural oils from Klaas and Warwel, 1999). Main epoxidation reaction by transfer-
agricultural resources, such as linseed oil and soybean oil, are ring oxygen from percarboxylic acid is always present with side
useful in polymer material syntheses. In order to produce a rigid reactions. During ring opening reaction, the side reactions are pos-
crosslinked thermoset from plant oils, the triglycerides must be sible due to percarboxylic acid which also plays role to produce
chemically functionalized by epoxidation with organic peracids epoxy groups, water, and hydrogen peroxide.
Catalytic processes are characterized by three phases: an acidic
solid catalyst phase, an aqueous phase which shows peracetic acid,
∗ Corresponding author. Tel.: +1 706 542 7353; fax: +1 706 542 4890. and the organic phase which contains epoxy groups (Campanella
E-mail address: ssharma@uga.edu (S. Sharma). and Baltanás, 2004). Diluting the organic third phase with a

0926-6690/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.indcrop.2011.10.036
486 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

Fig. 1. (A) Epoxidation, (B) side reactions (Rangarajan et al., 1995; Rüsch gen Klaas and Warwel, 1999), and (C) reaction between epoxy group and BPH (Kim et al., 2004).

solvent decreases the rate of side reaction that opens the epoxy ring. the unsaturation, the concentration of hydrogen peroxide and the
Espinoza Pérez et al. (2009) evaluated the effect of the solvent in the temperature simultaneously affected the percent and rate of oil
conversion to epoxy groups in plant oils. The solvent decreased the converted (Sinadinović-Fišer et al., 2001; Espinoza Pérez et al.,
rate of side reactions that opened the prior formed epoxy groups. 2009). Low levels of conversion indicated lower epoxy groups with
In solvent free process, hydrogen peroxide is extremely attractive low side reactions or lower epoxy groups with high side reactions.
and so it is the reagent used most. Hydrogen peroxide is more sol- High levels of conversion may show higher epoxy groups with
uble in organic solvent than in water, so that the production of low side reactions or higher epoxy groups with high side reac-
carboxylic acids, one type of side reaction, decreases (Vlček and tions. More dilute hydrogen peroxide promotes hydrolysis of the
Petrović, 2006). epoxy groups. Increasing the temperature can reduce the impact
Studies on epoxidation of different vegetable oils reported of the other factors; when the temperature decreases, the other
different experimental conditions catalyzed by an ion exchange factors become more significant. Hydrogen peroxide provides the
resin. Significant factors such as molar ratio of acetic acid and oxidant in the epoxidation process and is usually added in surplus
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 487

Table 1
The compositions of linseed, soybean, cottonseed, oilseed radish, and peanut oils.

Common name C:U (p) Fatty acid (%)


a
Linseed Soybean Oilseed radish Cottonseed Peanut Canola

Palmitic 16:0 4.65 10.36 6.13 19.09 9.49 4.3


Stearic 18:0 3.75 4.49 1.90 2.53 2.51 2.0
Oleic 18:1(1) 19.3 21.42 23.87 18.46 54.13 64.0
Linoleic 18:2(9,12) 15.13 53.54 13.46 55.47 23.80 18.4
Linolenic 18:3(9,12,15) 52.67 7.94 5.34 0.28 2.06 7.2
Arachidic 20:0 0.5 0.5 0.68 0.21 3.16 0.6
Gadoleic 20:1 0.12 0.37 8.58 0.12 0.17 1.1
Erucic 22:1(9) 0.09 0.4 31.76 0.16 0.72 NA

C, number of carbons of the fatty-acid chain; U, number of unsaturation; p, position of carbon after which the unsaturation is located.
a
Not tested, values as shown in Table 3 by Espinoza Pérez et al. (2009).

to ensure a complete reaction. It can also take part in degrad- its plastic forming ability using the least amount of the curing
ing the epoxy groups (Goud et al., 2007). An ion-exchange resin agent.
has been employed widely in epoxidation because it not only
exchanges other ions with materials in the solution without physi- 2. Materials and methods
cal alteration but also reduces the degradation of the epoxy groups
because of its large molecular size, which cannot allow it to go into 2.1. Materials
the resin after the peracetic acid has been produced (Gunstone,
1996). Several seed oils were purchased for this study: commercial-
Curing agents (hardeners) play an important role in determining grade linseed oil (Azure Farm, Inc., Dufur, OR, USA), cottonseed
the storage and handling requirements of epoxy formulations. Cur- oil (Glicks, Inc., Bloomfield, NJ, USA), soybean oil (Bakers & Chefs,
ing agents such as amines, anhydrides and dicyandiamide (DDA), Inc., Bentonville, AR, USA), and peanut oil (Hollywood Inc., Boul-
a latent curing agent, have been used as part of epoxy systems in der, CO, USA). The department of bioengineering at the University
adhesives, composites, printed circuit boards, and powder coatings of Georgia, Athens, GA, USA, provided oilseed radish oil. Vikoflex®
(Güthner and Hammer, 1993). Laliberte et al. (1983) found that (Gravity: 1.03, Oxirane oxygen: 9.18, iodine value: 3) was gifted
a one-part epoxy system consisting of DGEBA (diglycidyl ether of from Arkema Inc., Philadelphia, PA, USA. The following chemi-
bisphenol A) and DDA can be cured at a temperature of 180–200 ◦ C. cals were purchased from Sigma Chemical Co., St. Louis, MO, USA:
During polymerization or the curing reaction, epoxy groups open Amberlite IR 120H, acetic acid, H2 O2 , BF3 , pyrazine, benzyl bro-
due to the active hydrogen of the curing agents, creating a cross- mide, sodium hexafluoroantimonate, toluene, pyridine, n-butyl
linked network structure. Curing agents such as anhydrides and alcohol, ethanolic potassium hydroxide solution, phenolphthalein,
amines are available to cure epoxidized vegetable oils, but these and C17:0-heptadecanoic acid (>98% purity). All other materials
agents are toxic, require long curing processes at high tempera- used in this study were analytical-grade reagents.
tures, and have low heat resistance. Kim et al. (2004) reported
N-benzyl pyrazinium hexafluoroantimonate (BPH) to cure epoxy 2.2. Epoxidation of oils
resin systems were effective curing agents for various petroleum-
based epoxy resin systems as shown Fig. 1(C). Park et al. (2004a) Epoxidation of each oil was carried out in a round three-necked
used as little as 1% BPH to cure epoxy resin systems which exhibited 1 l flask equipped with a mechanical stirrer using a flexible Teflon
improved thermal-oxidative resistance and mechanical proper- blade 8 cm in diameter. A thermometer and an addition funnel
ties. They also showed that this curing agent did not propagate were attached in the flask. Epoxidation took place under the fol-
a reaction at a lower temperature (50 ◦ C), but created a rapid con- lowing conditions to yield high conversion and selectivity of each
version at a higher temperature (180 ◦ C), indicating that the agent oil (Espinoza Pérez et al., 2009): 1:1 molar ratio of CH3 COOH and
was latent at room temperature. Using this latent catalyst would unsaturation, 2:1 of H2 O2 and unsaturation, 30% (w/w) H2 O2 and
increase the storage stability and handling of thermosetting epoxy 25% resin by weight at 60 ◦ C without solvent. Each test used 150 g of
resins. linseed oil, cottonseed oil, soybean oil, oilseed radish oil, or peanut
The overall objective of this study was to develop environ- oil, as the case might be. The flask was placed in an oil bath and first
mentally sustainable bio-thermoset plastics and to compare the acetic acid and Amberlite IR 120H resin was added and stirred for
properties derived from linseed, soybean, cottonseed, oilseed 5 h at 500 rpm and 60 ◦ C, with gradual dropwise addition of H2 O2
radish, and peanut oils according to their different fatty acids. (2 ml/min). The yellow mixture was filtered through a cheesecloth.
Moreover, research was to investigate the thermal and ultrasonic The organic layer captured by the cheesecloth was washed with a
curing of natural, vegetable-based epoxy in order to determine saturated Na2 CO3 solution at 50 ◦ C and then dried using MgSO4 .

Table 2
Properties of oils after epoxidation.

Oil IV0 IVf OOe OOp Hydroxyl value (mg KOH/g) Conversion Selectivity

Linseed 189.41 31.02 8.91 11.07 7.3 0.81 0.97


Soybean 138.48 17.11 6.65 7.95 4.7 0.84 0.96
Radish 105.10 20.30 4.89 7.08 1.2 0.69 0.85
Cottonseed 117.93 8.12 5.52 6.93 1.8 0.80 0.86
Peanut 97.91 10.20 4.27 5.89 1.0 0.81 0.89
a
Canola 112.20 2.8 6.50 6.59 NA 98.5 1.00
a
Not tested, values as shown in treatment 13 in Table 4 by Espinoza Pérez et al. (2009).
488 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

Fig. 2. The composition of plant oils by gas chromatography (A), linseed oil (B), soybean oil (C), cottonseed (D), oilseed radish (E), peanut oil (*they were redrawn using Excel.
C17:0 as a standard).

2.3. Fatty-acid profile • Starting with 100 mg of oil in the glass tube, 20 mg ml−1
heptadecanoic acid (C17:0) was added as an internal
Before and after epoxidation, each oil was run on a gas chro- standard.
matograph using BF3 methylation following AOCS official method • Next, 2 ml of 0.5 N NaOH in methanol was added and then incu-
996.01, Sec. E (35). bated at 100 ◦ C for 5 min.
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 489

• Next, 2 ml of 14% BF3 , 2 ml of hexane, and 2 ml of saturated NaCl Injection (1 ␮l) was performed at a split ratio of 5:1. The carrier
were added. gas was helium supplied at a constant pressure, flowing at a rate
• The tube was vortexed and then centrifuged at 1000 rpm for of 1.1 ml/min. The injector temperature was 250 ◦ C and the FID set
10 min. point was 260 ◦ C.
• The upper layer was filtered with anhydrous sodium sulfate. The sample was held in the oven at 150 ◦ C for 3 min, followed by
an increase to 215 ◦ C at a pace of 10 ◦ C/min, and then held isother-
The gas chromatograph (Agilent Technology model 6890, mally for 40 min. The number of double bonds in a molecule (iodine
Wilmington, DE, USA) was equipped with a flame ionization detec- value) was determined for linseed, soybean, cottonseed, oilseed
tor (FID). Separation was achieved with an SP-2560 column (100 m, radish, and peanut oils using American Society for Testing and
0.25 mm i.d., and 0.20 ␮m film). Materials (ASTM) method D5554.

Fig. 3. The presence of epoxy groups and double bonds by FT-IR (A) epoxidized linseed (EL) and linseed and commercial Vikoflex® , (B) epoxidized soybean (ES) and soybean,
(C) epoxidized cottonseed (EC) and cottonseed, (D) epoxidized oilseed radish (ER) and oilseed radish, (E) epoxidized peanut (EP) and peanut oil.
490 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

Fig. 3. (continued)

2.4. Oxirane oxygen content and hydroxyl value of epoxidized oils • The white precipitate from the previous step was added to 20 ml
methanol and stirred at 70 ◦ C to dissolve all the precipitate and
The oxirane oxygen content and hydroxyl value of the then dried for several days to crystallize (Kim et al., 2004).
epoxidized oils were examined with American Oil Chemists’
Society (AOCS) official methods Cd 9-57 and Tx 1a-66, respec- 2.6. Preparation of plastics from epoxidized oils
tively.
BPH was dissolved in acetone, added to the epoxidized oils, vor-
texed for 5 min, and degassed for 60 min at room temperature. The
2.5. Synthesis of benzyl pyrazinium hexafluoroantimonate (BPH) mixture was poured into a Teflon mold, which was held in an oven
at 90 ◦ C for 1 h, and then at 140 ◦ C for 12 h.
• Pyrazine (5.49 g) was added to benzyl bromide (100.00 g) and
stirred for 48 h at 24 ◦ C under nitrogen gas to prevent additional 2.7. Analytical methods
reaction.
• After filtering, the precipitate was washed with toluene several Fourier Transform Infrared (FT-IR) spectra were recorded with
times and dried overnight under vacuum. 4 cm−1 resolution on an FT-IR spectrometer (Thermo Scientific,
• Sodium hexafluoroantimonate solution (14.99 g) was added to model Nicolet 6700, Waltham, MA, USA). Data was acquired using
the dried benzyl pyrazinium precipitate and stirred for 10 min. FT-IR software (Thermo Scientific, OMNIC series suite, Waltham,
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 491

Fig. 3. (continued)

MA, USA) which has been applied widely elsewhere for lipids, the unsaturation of the oils, whereas the oxirane oxygen content
because lipids have functional groups with absorption bands shows the presence of epoxy groups in the oils. Table 1 and Fig. 2
in the infrared region of the electromagnetic spectrum (Flatten show the initial iodine value and composition of peanut oil, oilseed
et al., 2005). 1 H nuclear magnetic resonance (NMR) spectra were radish oil, cottonseed oil, soybean oil, and linseed oils. Linseed oil
obtained on a spectrometer (Varian, Inc., model 300; Palo Alto, CA, had the highest iodine value because it contains the highest amount
USA) using chloroform (CDCl3 ) as a solvent. of ␣-linolenic acid, which has three double bonds in its fatty-acid
Thermal experiments were conducted using the following chain. Soybean and cottonseed oils, which each have two double
equipment to determine the extent of the resin conversions and bonds in their fatty-acid chain, showed around 54% and 55% linoleic
their thermal transitions into a thermoset matrix: the differen- acid, respectively.
tial scanning calorimeter (DSC) (Mettler Toledo, Inc., model 821; Conversion of double bonds to epoxy groups was followed by
Columbus, OH, USA), the thermogravimetric analyzer (TGA) (Met- an HBr titration. Table 2 shows a range of 4.9–8.9% oxirane oxy-
tler model 851) and the dynamic mechanical analyzer (DMA) gen (OOe ). Conversion ranged between 69% and 84%. The final
(Perkin-Elmer, model 8000; Waltham, MA, USA). The results of this iodine value after epoxidation showed numerous minor peaks;
testing helped to optimize the amount of curing hardener added to these peaks were not well resolved over 35 s of retention time
the resin. In addition, thermal analyses allow us to determine the because of the high molecular weight of oxygen in the fatty-acid
chemical changes that had occurred, such as oxidation and degra- chains. The low conversion rates were due to different initial iodine
dation. The curing behaviors of epoxidized oils were analyzed with values and reaction rates for linolenic acids, linoleic acids and oleic
the DSC at a heating rate of 10 ◦ C/min under N2 . Tests were per- acids (Scala and Wool, 2002). Linseed oil contains more than 50%
formed using the TGA to investigate the thermal degradation of linolenic acids in the triglycerides, whereas soybean and cottonseed
epoxidized oils from 25 ◦ C to 400 ◦ C at a heating rate of 10 ◦ C/min oils contain more than 50% linoleic acid.
under N2 . A scanning electron microscope (SEM) (Carl Zeiss SMT, Oleic, linoleic, and linolenic acids are comparatively complex:
Inc., model 1450; Peabody, MA, USA) was used to investigate the while each is 18 carbon atoms long, they each have a different num-
morphology of the bioplastics produced. ber of double bonds and their double bonds appear at different
Oven-cured specimens were tested on the DMA for their locations along their fatty-acid chains. In contrast, oilseed radish
mechanical performance over a temperature range of −30 ◦ C to oil is relatively simple: it has 32% erucic acid containing one double
200 ◦ C at a frequency of 1.0 Hz. The heating rate was 2 ◦ C/min and bond located at the 13th position along its fatty-acid chain from the
amplitude was 15 ␮m. A three-point bending test was used to glycerol center.
determine the damping performance of each bioplastic by eval- There are at least two possible explanations for the variation
uating its storage modulus, loss modulus and glass transition in epoxidation results among the oils in this study. The first possi-
temperature (O’Donnell et al., 2004). bility is that the differences arise from steric factors such as that
double bonds relatively distant from the glycerol center of the
fatty-acid chain are more reactive than double bonds near the
3. Results and discussion
glycerol center. The second possibility looks to electronic effects.
3.1. Iodine value and oxirane oxygen percent Free fatty-acids have a higher reaction rate of epoxidation than the
triglycerides. As the number of double bonds increases, the electron
The iodine value and the oxirane oxygen content are important density also increases, which causes the rate constant to increase
characteristics in the epoxidation of oils. The iodine value specifies (Gunstone, 1996). In particular, linseed oil, which has a high level
492 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

of linolenic acids, reported partial epoxidation with side reactions by solutions in the process. Epoxidation is a simultaneous process
(Chen et al., 2002; Zong et al., 2005). In contrast, the erucic acids in that produces both epoxy and hydroxyl groups from side reactions
oilseed radish oil, containing a mono-double-bond fatty-acid chain with H2 O2 , H2 O, and CH3 COOOH, as shown schematically in Fig. 1.
22 hydrocarbons long, showed a low conversion ratio. In this study, the higher initial iodine value contributed to higher
The range of selectivity was mostly high, around 0.85–0.97. The epoxy conversion and also to higher hydroxyl values.
highest selectivity was for linseed oil (around 0.97) due to having Fatty acid chain composition has a significant effect on the rate
more double bonds. In the hydroxyl value test, epoxidized linseed of epoxidation. The double bonds in oleic and linoleic acid were
oil (EL) and epoxidized soybean oil (ES) showed higher readings (7.3 equally reactive and the double bonds of linolenic acid were about
and 4.7, respectively) than epoxidized cottonseed oil (EC), epoxi- two times more reactive than oleic and linoleic acids (Scala and
dized oilseed radish (ER), or epoxidized peanut oil (EP), under 1.8 Wool, 2002). Erucic acid (22:1, cis-13-docosoenoic acid) is present
because EL and ES have more epoxy groups that can be attacked at high levels in seed oils. It can also be oxidized at the 13th atom of

Fig. 4. The partially epoxidized plant oils by 1 H NMR (A) EL, (B) ES, (C) EC, (D) ER, and (E) EP.
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 493

Fig. 4. (continued)

carbon from the glycerol center. A hydrocarbon chain, in its most the organic phase of the epoxidation processes (Johnson and Fritz,
stable zigzag form, is characterized by dense, tightly packed chains. 1989). The long fatty-acid chain in erucic acid is less soluble than
Long-chain fatty-acid chains bend slightly more than a medium oleic acid in the organic phase. It may, therefore, induce a weaker
fatty-acid chain (Markley and Klare, 1947). Shorter chains spread attack by hydrogen peroxide or a stronger attack by peracetic acid
on water, but longer chains form a lens, indicating that the oil’s or water, in either case producing less conversion. We are left to
work of adhesion to water is less than the work of cohesion (Small conclude that the different reaction rates among the oils used in
et al., 1986). Additionally, short and medium fatty-acid chains are this study were caused by intrinsic properties such as composition
100 times more water soluble than the long-chain fatty-acids, due of the oils, since the solubility of external solutions or and other
to their small molecular size, so they showed different affinity in extrinsic conditions are similar.
494 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

Fig. 4. (continued)

3.2. The double bonds and epoxy groups characterized by FT-IR, soybean, cottonseed, oilseed radish, and peanut oils showed a simi-
1 H NMR lar pattern, but ER and EP showed low intensity of absorbance in the
FT-IR because the percent of oxirane oxygen was low (around 4.9%
The bands displayed by the FT-IR at 3010 cm−1 were attributed and 4.3%, respectively), attributable to fewer epoxy groups than in
to C–H stretching of C C–H, and the very weak minor bands at other oils such as EL, ES, and EC.
1651.0 cm−1 and 3052.0 cm−1 were due to CH CH stretching. The Fig. 4 shows the 1 H NMR spectra of oils with different degrees of
intensity of the 3010.0 cm−1 band decreased with the increasing epoxidation. New resonances at ı = 2.97–3.13 ppm were identified
degree of epoxidation (Chen et al., 2002). A new band appeared which were attributed to H 2 , H 3 , H 4 , and H 10 for ring formation
in the spectra of epoxidized oils at 823.0 cm−1 attributed to epoxy of epoxide (Zong et al., 2005; Chen et al., 2003). The intensity of the
groups. The intensity of this band increased with the increasing resonance at ı = 2.97–3.13 ppm increased with increasing epoxida-
degree of epoxidation. Fig. 3(A) indicates the presence of unsat- tion, which represents the CH protons attached to the oxygen atoms
uration in the linseed oil at 3010.0 cm−1 prior to epoxidation. of both the epoxy group, whereas the double bonds are indicated
Epoxidized linseed oil and Vikoflex® linseed oil showed no peak in at 5.2–5.5 ppm. The resonance between 3.8 and 4.4 indicated the
this region, but displayed the presence of epoxy groups in the epoxi- glycerol center in the triglycerides (Scala and Wool, 2002; Chen
dized linseed and Vikoflex® epoxy resins at 823.8 cm−1 . Epoxidized et al., 2002). 1 H NMR allows the researcher to detect the presence

Fig. 5. TGA thermographs of the as-received natural and after epoxide of linseed, soybean, cottonseed, oilseed radish, and peanut oils. Methods: Thermal scan was carried
on a TGA instrument at a heating rate of 20 ◦ C/min to observe the thermal stability of these functionalized oils.
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 495

Fig. 6. The heat flow analysis of epoxidized oils with 1% weight of BPH, the curing agent by DSC, M.P.: melting point.
496 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

Fig. 7. Three points bending test of bio-thermoset plastics from epoxidized oils by dynamic mechanical analysis.

of hydroxyl groups attached to the triglycerides at 3.8 ppm (Farias chains result in relatively high melting points (Markley and Klare,
et al., 2010). EL and ES portrayed a high concentration of hydroxyl 1947). Linolenic acid, with its three double bonds, has been used
groups in the graphs. Table 1 shows higher hydroxyl values for EL as an agent to increase drying and film-forming properties in prod-
and ES; but EC, ES, and EP have lower hydroxyl values. EL and ucts manufactured by the paint, varnish and linoleum industries.
ES have a greater number of double bonds and epoxide groups Introducing one or more double bonds into the hydrocarbon chain
attacked by H2 O2 , H2 O, and CH3 COOOH. The values in Table 1 are of unsaturated fatty-acids results in one or more “bends” in the
also due to simultaneous epoxidation and side reactions in the molecule. The geometry of the double bond is almost always a cis
epoxidation process. configuration in natural fatty-acids. These molecules do not “stack”
very well; the intermolecular interactions are much weaker than
3.3. Thermal characterization of vegetable oil resins and their in saturated molecules. As a result, the melting points are much
thermoset plastics lower for unsaturated fatty-acids. Fig. 6 shows the melting points
of the oils in this study. Linseed oil, with the highest composition
3.3.1. Thermal stability of oils before and after functionalization of three double bonds on fatty-acids, has a melting point of −40 ◦ C.
Fig. 5 indicates that as measured by the TGA, the vegetable oils Soybean oil, cottonseed oil, oilseed radish oil, and peanut oil have
showed good thermal stability before and after epoxidation. While melting points of −20 ◦ C, −8 ◦ C, 4 ◦ C, and 4 ◦ C, respectively. Dou-
most of the oils begin degrading at approximately 400 ◦ C, soybean ble bonds result in bends in the molecule which makes it harder to
and linseed oils start to lose weight from 150 ◦ C onward. After epox- ‘pack’ these molecules into a low-energy crystalline morphology.
idation, linseed oil and soybean oils started degrading at higher Cottonseed, oilseed radish, and peanut oils show a stable line after
temperature than before epoxidation. As shown in Fig. 6(A and B), melting point, but linseed and soybean oils indicated decreasing
linseed and soybean oils lost their thermal stability after reach- heat flow over 150 ◦ C.
ing their melting point; however, Fig. 6(C–E) shows a straight line
after reaching their melting point. The formation of an epoxy group 3.3.2. The thermal curing of epoxidized oils
increased thermal stability as shown in Fig. 5(B). We also conducted an isothermal curing study of all epoxi-
In general, the unsaturated fatty-acids have lower melting dized vegetable oils in the differential scanning calorimeter (DSC)
points than the saturated fatty-acids. The additional double bonds at an isothermal temperature of 140 ◦ C. The typical thermograph
depress the melting point for oleic, linoleic, and linolenic to of isothermal curing from the DSC instrument is comprised of an
13–16 ◦ C, −5 ◦ C, and −11 ◦ C, respectively, whereas close inter- initial baseline, an exothermic peak, and a final baseline. When the
molecular interactions such as longer and saturated fatty-acid initial baseline and final baseline are at the same level, the curing

Table 3
Crosslinking density of epoxy resin.

Resin with 1% BPH Tg (K) E (MPa)  (mol/cm3 )

EL 327.3 170.1 6.27 × 10−2


ES 326.3 161.3 5.92 × 10−2
EC 320.5 101.0 3.79 × 10−2
ER 300.6 97.4 3.67 × 10−2
EP 319.3 68.2 2.73 × 10−2
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 497

Fig. 8. The morphology of bioplastics from epoxidized oils with 1% BPH after curing by scanning electron microscopy from (A) EL, (B) ES, (C) EC, (D) ER, and (E) EP.
498 J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499

is assumed to be complete. The rate of reaction is proportional to Tg . Variation in initial storage modulus contributed to different oils
the evolved rate of heat in DSC analysis. Therefore, the reaction depended upon the number of epoxy groups that react with BPH.
rate (dt/dc) was determined for different degrees of curing at these Higher epoxy groups give tough materials because of more reac-
isothermal temperatures. As shown in Fig. 6, these graphs clearly tions with BPH. Oil-based plastics showed a significant amount
show the autocatalytic behavior of epoxy curing. In addition, the of the modulus above the glass transition region in the rubbery
exothermic curves consist of an initial shoulder (initiation) and plateau. The plastic from EP showed the lowest tan ı and storage
main exothermic peaks (curing reaction). The polymerization of modulus values, but lost storage modulus at temperatures below
a cationic epoxy system leads to a Lewis acid process with two 50 ◦ C. The plastics from EL, ES, and EC indicated similar tan ı over
separate initiation reactions (Boquillon and Fringant, 2000). The 50 ◦ C, but they showed a different initial storage modulus. The plas-
polymerization of a vegetable-oil-based cationic epoxy system sim- tic from EL promised the highest modulus due to larger number of
ilarly leads to a Lewis acid process with two separate initiation epoxy groups in the triglycerides and fatty-acid chains as evident
reactions (Bailey, 1964): from the higher oxirane oxygen number. The plastics from EP and

ER shows lower toughness similar to rubbery materials that it can
BPH + Epoxide−→Ph–CH2 –(O+ of epoxide)–SbF−
6
+ pyrazine be used in the form of elastomeric polyurethane for shoe soles and
(1) insulation (Pothan et al., 2003). We observed that the plastic from
EP, ER and EC lacked integrity and crumbled compared to EL and
ES plastics, which demonstrated good integrity and showed the
 dynamic mechanical characteristics of a rubbery material.
BPH + R-OH−→Ph–CH2 –OR + pyrazine + H+ SbF−
6
(2)
DMA analysis provides information about the viscoelasticity,
The two reactions shown in chemical reactions (1) and (2) are crosslinking density and thermal stability of polymer networks.
responsible for the initial shoulder of the exothermic curves, indi- Crosslinking density can be obtained from the following equation
cating the initiation of the curing reaction, and the subsequent main for rubber elasticity in a plateau region of the DMA curve (Park et al.,
exothermic peak reflecting the catalytic action of HSbF6 . It was 2004b).
observed that 1% of BPH was sufficient to cure epoxidized oils fully.
BPH exhibited a sharp endothermic peak at a temperature of 120 ◦ C, E
=
resulting in its dissociation, which activates the epoxy curing reac- ˚RT
tion. Integrating the thermographs disclosed that the exothermic
heat of curing was 70.55 J/g for EL, 59.34 J/g for ES, 49.28 J/g for EC, where E represents the rubber modulus,  the density of network
42.81 J/g for ER, and 36.66 J/g for the EP system. These differences or crosslinking density (mol/cm3 ), ˚ the front factor (usually equal
in the heat of curing among the vegetable oils can be attributed to unity), R the gas constant, and T the absolute temperature in the
to EL having a larger number of epoxy groups than other oils due rubbery region.
to the greater degree of unsaturation from constituent esterified Table 3 shows the storage modulus in the rubbery plateau and
linolenic acid, which contains three double bonds. Fig. 6 shows the the glass transition temperature of the epoxidized oils system-
DSC thermograph for epoxidized oils during and after isothermal cured using 1% by weight of BPH. The rubbery storage modulus
curing, in the absence of the curing agent. The presence of the epoxy resulted in different values across the tested oil-based epoxy resins,
rings inhibits close packing the fatty-acid chains, preventing the while the glass transition temperatures were similar. This differ-
transition of the liquid state to the grease state (Wool and Sun, ence in the heat of curing can be attributed to the higher number
2005). of epoxy groups in epoxidized oils due to the higher degree of
unsaturation from the constituent esterfied linolenic acid, which
3.4. The performance of bio-thermoset plastics contains three double bonds. As seen in Table 3, crosslinking density
increases as epoxy groups are created. Adding the heavier molecule
After curing using BPH, we conducted DMA analysis using the weight of epoxy groups into oils increases the impact-resistance
thermal three-points bending mode at temperatures ranging from property of the epoxy, resulting in tougher composites. The results
−30 ◦ C to 200 ◦ C. Dynamic mechanical properties depend on the showed that the epoxidized oils were superior to the other rubbers
properties of component polymers. The glass transition tempera- as an impact modifier in both thermal properties and environmen-
ture gives an indication about the density of cross-linked polymers. tal compatibility (Park et al., 2004a). Overall, this thermal stability
A formulation containing a lower percentage of the curing agent would produce minimum volatiles and consequently less shrinkage
showed a single relaxation located at a lower temperature. That during the thermal molding process.
means that epoxidized oils are partially miscible with the epoxy The morphological microstructures from scanning electron
resins, decreasing the cross-linking density of the sample (Rebizant microscopy were used to examine the fracture properties of all
et al., 2003). neat epoxidized resins. Fig. 8 shows the morphologies of bioplas-
Tan ı relates the impact resistance of a material and the ratio of tics from studied epoxidized oils. The bioplastics from EL, ES and
the loss modulus to the storage modulus. The peaks in tan ı occur in EC demonstrated rippled or fiber-like (Miyagawa et al., 2004) frac-
the region of the glass transition where the material changes from ture morphologies indicating their good toughness characteristics
a rigid state to a more elastic state; the measure is associated with which may be attributed to higher epoxidation conversion and
the movement of small groups and chains of molecules within the selectivity due to presence of more linoleic and linolenic fatty-
polymer structure. Glass transition temperature (Tg) represents the acid chains. This phenomenon of toughness is more prominent in
single peak of the tan ı attributed to a higher value with increasing EL than ES and less in EC neat plastics. The bioplastics from EP
content of iodine value and oxirane oxygen percent. The higher the and ER exhibited more fragmentations (powdered particles around
tan ı peak value, the greater is the degree of molecular mobility surface) on fracture indicating lack of structural integrity which
(Kuzak and Shanmugam, 1999). is also evident from their lower dynamic storage modulus. This
Fig. 7 shows the effect of temperature on the dynamic mechan- phenomenon indicates lack of unsaturation, especially due to the
ical performance of bioplastics from various epoxidized oils. All fatty-acid chains such as linoleic and linolenic. Uniform distribu-
samples exhibited a drop in storage modulus as temperature tion of the matrix was observed very important for toughening,
increased because they became more mobile and moved into a because it allowed the yielding process to operate throughout the
wider packing arrangement in rubbery plateau area approaching matrix (Ratna, 2001). Therefore, the neat resin matrix from EL, ES
J.R. Kim, S. Sharma / Industrial Crops and Products 36 (2012) 485–499 499

and EC showed more uniform fracture surface, indicating greater Flatten, A., Bryhni, E.A., Kohler, A., Egelandsdal, B., Isaksson, T., 2005. Determination
toughness and impact strength than EP and ER bioplastics. of C22:5 and C22:6 marine fatty acids in pork fat with Fourier transform mid-
infrared spectroscopy. Meat sci. 69 (3), 433–440.
Goud, V., Pradhan, N., Dinda, S., Pradhan, N.C., 2007. Epoxidation of Karanja
4. Conclusions (Pongamia glabra) oil catalysed by acidic ion exchange resin. Eur. J. Lipid Sci.
Tech. 109 (6), 575–584.
Guenter, S., Rieth, R., Rowbotton, K.T., 2003. Ullmann’s Encyclopedia of Industrial
In this study, an epoxidation by hydrogen peroxide along with a Chemistry, vol. 12., 6th ed. Wiley, New York, pp. 269–284.
catalyst was the most efficient method to epoxidize oils promising Gunstone, Frank, 1996. Fatty-acid and Lipid Chemistry, 1st ed. Springer-Verlag.
a high conversion percentage. Fatty epoxides can be used directly Güthner, T., Hammer, B., 1993. Curing of epoxy resins with dicyandiamide and
urones. J. Appl. Polym. 50, 1453–1490.
as plasticizers to improve the flexibility, elasticity and stability of Johnson, R., Fritz, E., 1989. Fatty-acids in Industry. Marcel Dekker Inc., NY, USA.
polymer subjected to heat. The solvent-free epoxidation process Kim, M.S., Lee, K.W., Endo, T., Lee, S.B., 2004. Benzylpyrazinium salts as thermally
generally showed suitable oxirane oxygen levels and less oxirane latent initiators in the polymerization of glycidyl phenyl ether: substituent
effect on the initiator activity and mechanistic aspects. Macromolecules 37,
degradation. In this study, adding 1% of BPH by weight as a cur- 5830–5834.
ing agent was sufficient to fully cure epoxidized oils and contribute Kuzak, S.G., Shanmugam, A., 1999. Dynamic mechanical analysis of fiber-reinforced
light weight and good thermal stability. The chain length of fatty- phenolics. J. Appl. Polym. Sci. 73, 649–658.
Laliberte, B.R., Bornstein, J., Sacher, R.E., 1983. Cure behavior of an epoxy resin-
acids and the number or locations of double bonds give different
dicyandiamide system accelerated by monuron. Ind. Eng. Chem. Prod. Res. 22,
properties to the bio-thermoset plastics in this study. Linseed oil 261–262.
containing linolenic acids promised the highest modulus and good Markley, Klare, S., 1947. Fatty-acids: Their Chemistry and Physical Properties. Inter-
impact resistance properties. Shorter hydrocarbon length on fatty- science Pub. Inc., NY, USA.
Miyagawa, H., Mohanty, A.K., Misra, M., Drzal, L.T., 2004. Thermo-physical and
acid chains and the number of double bonds (a higher iodine value) impact properties of epoxy containing epoxidized linseed oil. Macromol. Mater.
enabled higher selectivity and conversion of epoxy groups to con- Eng. 289, 629–635.
fer tougher plastics matrices. Future work may focus on studying Mungroo, R., Pradhan, N., Goud, V., Dalai, A., 2008. Epoxidation of canola oil with
hydrogen peroxide catalyzed by acidic ion exchange resin. J. Am. Oil Chem. Soc.
fiber-reinforced composites created by adding crystalline natural 85, 887–896.
fiber and blends of vegetable oil-based epoxy resins to improve the O’Donnell, A., Dweib, M., Wool, R., 2004. Natural fiber composites with plant oil-
strength of composites while providing complete biodegradability based resin. Compos. Sci. Technol. 64, 1135–1145.
Park, S.J., Jin, F.L., Lee, J.R., 2004a. Thermal and mechanical properties of tetrafunc-
and appropriate quality. tional epoxy resin toughened with epoxidized soybean oil. Mater. Sci. Eng. A
374., 109–114.
Acknowledgements Park, S.J., Lee, H.Y., Han, M., Hong, S.K., 2004b. Thermal and mechanical interfa-
cial properties of the DGEBA/PMR-15 blend system. J. Colloid Interface Sci. 270,
288–294.
Portions of this research were supported by an award Pothan, L.A., Oommen, Z., Thomas, S., 2003. Dynamic mechanical analysis of banana
(BSRI1102) from the Bioenergy Systems Research Initiative at the fiber reinforced polyester composites. Compos. Sci. Technol. 63, 283–293.
Rangarajan, B., Havey, A., Grulke, E.A., Culnan, P.D., 1995. Kinetics of the parameters
University of Georgia. We acknowledge Dr. Casimir C. Akoh for of a two-phase model for in situ epoxidation of soybean oil. J. Am. Oil Chem. Soc.
allowing us to use his GC equipment and Dr. Vikram Dhende for 72 (10), 1161–1169.
synthesis of BPH. Ratna, D., 2001. Mechanical properties and morphology of epoxidized soyabean-oil-
modified epoxy resin. Polym. Int. 50, 179–184.
Rebizant, V., Abetz, V., Tournilhac, F., Court, F., Leibler, L., 2003. Reactive tetrablock
References copolymers containing glycidyl methacrylate. Synthesis and morphology con-
trol in epoxy–amine networks. Macromolecules 36 (26), 9889–9896.
Adachi, T., Araki, W., Nakahara, T., Yamaji, A., Gamou, M., 2002. Fracture toughness Rüsch gen Klaas, M., Warwel, S., 1999. Complete and partial epoxidation of plant
of silica particulate-filled epoxy composite. J. Appl. Polym. Sci. 86, 2261. oils by lipase-catalyzed perhydrolysis. Ind. Crops Prod. 9, 125–132.
Bailey, A.E., 1964. Industrial Oil and Fat Products, 2nd ed. The Interscience Publishers Scala, J., Wool, R., 2002. Effect of FA composition on epoxidation kinetic of TAG. J.
Inc., NY, USA. Am. Oil Chem. Soc. 79 (4), 373–378.
Boquillon, N., Fringant, C., 2000. Polymer networks derived from curing of epoxidised Sinadinović-Fišer, S., Janković, M., Petrović, Z.S., 2001. Kinetics of in situ of epoxida-
linseed oil: influence of different catalysts and anhydride hardeners. Polymer 41, tion of soybean oil in bulk catalyzed by ion exchange resin. J. Am. Oil Chem. Soc.
8603–8613. 78 (7), 725–731.
Campanella, A., Baltanás, M.A., 2004. Degradation of the oxirane ring of epoxidized Small, D., Craven, B., Yvonne, L., Shipley, G., Steiner, J., 1986. Handbook of Lipid
vegetable oils with hydrogen peroxide using an ion exchange resin. Eur. J. Lipid Research; the Physical Chemistry of Lipids, vol. 4. Plenum Press, NY, USA.
Tech. 106, 524–530. Vlček, T., Petrović, Z.S., 2006. Optimization of the chemoenzymatic epoxidation of
Can, E., Wool, R.P., Küsefoğlu, S., 2006. Soybean- and castor-oil-based thermosetting soybean oil. J. Am. Oil Chem. Soc. 83, 247–252.
polymers: mechanical properties. J. Appl. Polym. Sci. 102, 1497–1504. Williamson, A., 2001. Encyclopedia of Materials Science & Technology. Elsevier Sci
Chen, J., Soucek, M.D., Simonsick, W., Celikay, R., 2002. Epoxidation of partially Ltd, p. 90.
norbornylized linseed oil. Macromol. Chem. Phys. 203 (14), 2042–2057. Wool, R.P., Sun, X.S., 2005. Bio-based Polymer and Composites, 1st ed. Academic
Dinda, S., Goud, V., Patwardhan, A.V., Pradhan, V., 2008. Epoxidation of cottonseed oil Press Inc., NY, USA.
by aqueous hydrogen peroxide catalyzed by liquid inorganic acids. J. Bioresour. Zaher, F.A., El-Mallah, M., El-Hefnawy, M., 1989. Kinetics of oxirane cleavage in
Technol. 99 (9), 3737–3744. epoxidized soybean oil. J. Am. Oil Chem. Soc. 66, 698–700.
Espinoza Pérez, J.D., Haagenson, D.M., Pryor, S., Ulven, C., Wiesenborn, D., 2009. Pro- Zhao, H., Zhang, J., Sun, X.S., Hua, D.H., 2008. Syntheses and properties of cross-
duction and characterization of epoxidized canola oil. ASABE 52 (4), 1289–1297. linked polymers from functionalized triglycerides. J. Appl. Polym. Sci. 110,
Farias, M., Martinelli, M., Bottega, D.P., 2010. Epoxidation of soybean oil using a 647–656.
homogeneous catalytic system based on a molybdenum (VI) complex. Appl. Zong, Z., He, J., Soucek, M.D., 2005. UV-curable organic–inorganic hybrid films based
Catal. A: Gen. 384, 213–219. on epoxynorbornene linseed oils. Prog. Org. Coat. 53 (2), 83–90.

You might also like