+ All Categories
Home > Documents > Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

Date post: 14-Apr-2018
Category:
Upload: victor-lopez
View: 230 times
Download: 0 times
Share this document with a friend

of 15

Transcript
  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    1/15

    1

    Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent for

    Epoxy Resin

    Rongpeng Wang and Thomas Schuman

    Department of Chemistry, Missouri University of Science and Technology, Rolla, MO 65409

    Abstract

    Epoxidized glycidyl esters of soybean oil (EGS) have been synthesized and used as reactive diluents for partial

    replacement of a commercial, bisphenol A-based epoxy resin (DGEBA). The EGS merits include a higher epoxy

    content and lower viscosity than the epoxidized triglyceride soybean oil (ESO). Thermosetting resins were

    fabricated from DGEBA systems blended with various amounts of EGS and ESO, using 4-methyl-1,2-

    cyclohexanedicarboxylic anhydride as a curing agent and 2-ethyl-4-methylimidazole as catalyst. The curing

    behavior and glass transition were monitored by differential scanning calorimetry (DSC), the performance of

    thermosetting resins was studied by measurement of thermal stability and flexural properties. The results

    indicate that EGS resins provide better compatibility, intermolecular crosslinking, and yield materials that are

    stronger than materials obtained using ESO. However, the EGS resin systems significantly reduce viscosity

    compared to either pure DGEBA or ESO-blended DGEBA counterparts. Therefore, EGS derived from renewable

    sources holds potential to enable fabrication of complex, shaped epoxy composites for structural applications.

    1. IntroductionEpoxy resin is one of the most important thermosetting polymers and widely used in coatings, adhesives

    and composites due to its excellent mechanical strength, outstanding chemical resistance, good thermal and

    electrical properties, and low shrinkage upon cure.1 Most commercially epoxy resins are relatively high

    viscosity liquids or solids. Cured pure epoxy resins are rigid and brittle materials with low impact resistance.

    Adding linear elastomeric or thermoplastic additives can increase the toughness; however, this invariably

    results in a corresponding decrease in resin flow and processing difficulty.2

    Reactive diluents are used for reducing and controlling the viscosity of epoxy resins to improve wetting

    and handling characteristics because in the liquid-moulding technologies like resin transfer moulding or

    pultrusion, the viscosity and resin flow are critical to achieving a quality laminate.3 Recent trends toward lower

    Correspondence to: T. Schuman ([email protected])

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    2/15

    2

    VOC, higher solids epoxy formulations have also resulted in increased utilization of reactive diluents.1 Phenyl

    glycidyl ether and n-butyl glycidyl ether are efficient and widely used diluents, but they have been losing

    interest due to toxicity, volatility or obnoxious odor issues. Industrial trends show increasing interest in longer

    chain reactive diluents, e.g., C12C14 alkyl glycidyl ethers, neopentyl glycol diglycidyl ether, diglycidyl ether of

    polypropylene glycol. These longer chain reactive diluents can also behave as flexibilizing agents to increase

    thermosetting polymer elongation and impact resistance, though there is often a tradeoff of tensile strength,

    glass transition temperature and chemical resistance.1

    Since petroleum resources are limited, polymers based on vegetable oils are of great interest because they

    are renewable and can significantly contribute to a more sustainable development.4-6

    Epoxidized soybean oil

    (ESO) has attracted great interest because of a plentiful soybean supply in the United States and therefore of

    relatively low cost. ESO can be crosslinked into thermosetting polymers by various curing agents.7

    However,

    due to a low oxirane content and sluggish reactivity of internal epoxy groups, the cured ESO normally has a lowcrosslinking density, resulting from partially unreacted ESO and saturated fatty acid chains that act to plasticize

    and degrade the thermal and mechanical properties of cured resin.

    Most ESO industrial uses are limited to nonstructural applications such as plasticizers or stabilizers for poly

    vinyl chloride (PVC)8, oil-base coatings

    9with low strength requirements.

    10ESO has a moderate viscosity and

    offers good miscibility with epoxy resins.11 So ESO or their derivatives can be used as reactive diluents for the

    partial replacement of epoxy resin to decrease the overall cost and improve the processability.2,12,13 Generally,

    the mechanical strengths of ESO blended resins are not comparable to those of pure, non-modified epoxy

    resins, while their toughnesses are better due to the introduction of a two phase structure.14-17

    Only those oils of poly-unsaturated fatty acid content, especially soybean or linseed oils that can produce

    dense epoxy functionality resins, are capable to produce satisfactory properties.7,18,19 Epoxidized vegetable oils

    (EVO) of low oxirane values either are not reactive or impart waxy, non-curing properties to the resin system. In

    this research, epoxy resins of epoxidized glycidyl ester (EGS) derived from soybean oils were synthesized and

    examined. The goals were to remove as much as possible the plasticizing effect of saturated components, to

    further increase the oxirane content, and to minimize viscosity to permit use as an efficient reactive diluent.

    We reason that EGS with the addition of a terminal epoxy group (glycidyl), which is readily accessible to

    nucleophilic attack, should further support reactivity compared with standard, commercial ESO. The increase

    in oxirane content, consequent reduction of the ESO molecular size, and a facilitated removal of the saturated

    component should each provide a denser intermolecular crosslinking structure and yield a thermosetting resin

    material with improved properties.

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    3/15

    3

    2. Experimental2.1. Materials

    Refined, food grade soybean oil was purchased from a local grocery store. ESO was purchased from Union

    Carbide Corporation. Epichlorohydrin, methylene chloride, meta-chloroperoxybenzoic acid (MCPBA), sodium

    carbonate, sodium bicarbonate, sodium sulfite, and anhydrous sodium sulfate were purchased from Fisher

    Scientific. Cetyltriethylammonium bromide (CTEAB), 2-ethyl-4-methylimidazole, and 4-methyl-1,2-

    cyclohexanedicarboxylic anhydride (MHHPA) were purchased from Aldrich. Commercial DGEBA was supplied

    by Momentive with trade name EPON Resin 828. Mold release agent Chemlease 41-90 EZ was purchased

    from Chem-Trend, Inc.

    2.2. Soap and free fatty acid preparationSoybean oil derived free fatty acids (FFA) were made via acid neutralization of soap. Vegetable oil was

    reacted with sodium hydroxide to generate soap, then acidified with sulfuric acid to pH

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    4/15

    4

    Glycidyl ester (GE) and sodium carbonate were mixed with methylene chloride. MCPBA dissolved in

    methylene chloride was added dropwise at a reaction temperature below 15C for 4 hours to complete

    epoxidation. The reaction mixture was washed with 10% sodium sulfite and then by 10% sodium bicarbonate

    and water. Methylene chloride was removed by in vacuo rotary evaporation and the product EGS was dried

    using anhydrous sodium sulfate. Epoxidation also was accomplished by an analogous performic acid process

    except using toluene as solvent instead of benzene.20

    2.4. CharacterizationInfrared spectra (IR) were conducted on a Nicolet Nexus 470 E.S.P. spectrophotometer.

    1H NMR spectra

    were obtained on a Varian VXR 400 MHz spectrometer using DMSO-d6 as solvent. Iodine value measurements

    were based on ASTM Method D554-95. Oxirane value was measured using AOCS Method Cd 9-57. Viscosity

    was tested on a Brookfield LVDV-III+ Ultra Rheometer at 25C.

    2.5. Curing reactionThe weight ratios of EGS/ESO to DGEBA resin blend chosen for the present work were 0:100 (pure DGEBA),

    10:90, 30:70, 50:50, 70:30; and 90:10. Stoichiometric weight of MHHPA curing agent and 1 wt% (based on

    epoxy part) of 2-ethyl-4-methylimidazole were added to the epoxy resin blend. After mixing by a high shear

    mixer for 10 min, the mixture was degassed for 30 min, then poured into a mold sprayed with a mold release

    agent. Curing was performed at 145o

    C for 15 hrs for all blends except ESO-DGEBA (90:10) blend, which was

    firstly pre-cured at 145oC for 10 min then mixed again and poured into the mould for fully cure at 155

    oC for 15

    hrs, because of the low reactivity of ESO. The postcure for all samples were performed at 175oC for 1 hr.

    The mixtures weighing 2 to 3 mg encapsulated in aluminum hermetically sealed pan were also cured on a

    differential scanning calorimetry machine at a heating rate of 10 oC/min from 40-250 oC to study the cure

    behavior of the different formulations.

    2.6. Thermal TestsDSC (Q2000, TA Instrument) was used to determine the glass transition temperature (Tg) of cured resin.

    Measurement was carried out over a temperature range from -40 to 180 C at a heating rate of 20 C/min.

    Samples were first preheated to 180 C and quenched with liquid N2 to remove any thermal history.

    TGA (Q50, TA Instrument) was used to determine the thermal stability of cured resin. Measurement was

    performed from 30 to 750 C at a heating rate of 10 C/min under an ambient air flow environment.

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    5/15

    5

    2.7. Mechanical TestsThe flexural properties were determined according to the ASTM method D790 on an Instron 4469

    universal testing machine. The modulus was determined in a three-point bending mode, with a sample

    dimension of 102 mm12.7 mm3.2 mm. The span was 50.8 mm, the crosshead speed was set at 12.7

    mm/min.

    3. Results and Discussion3.1. Preparation of Epoxidized Glycidyl Ester

    Scheme 1. Synthetic route to EGS. (Soybean oil and fatty acids are shown as simplified structures

    containing only oleic acid though they also contain saturated and polyunsaturated fatty acids. See the text for

    detail.)

    Scheme 1 shows the synthetic route to EGS, oleic acid generalizing a soybean fatty acid chain. Preparation

    of mixed FFA from triglyceride is straightforward and well-developed. Acetone was used as a low boiling

    recoverable solvent to prepare soap and effect low temperature crystallization. A slight excess of NaOH with

    higher concentration was desirable when preparing soap from FFA because unsaturated FFAs are prone to

    dissolve in acetone rather than react with base and also unsaturated FFA soap is more soluble in water.21

    Carefully dried and finely powdered soap resulted in greater yields of glycidyl esters of fatty acids.22

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    6/15

    6

    A low solubility of soap in epichlorohydrin suggested a phase transfer catalyst was needed to accelerate

    the reaction. With CTEAB catalyst, the consumption of soap was completed within half an hour under reflux

    condition. Glycidyl esters can also be prepared directly from FFA in EPCH rich medium, and then subsequently

    dehydrohalogenating with alkali, but the yield and purity were lower than the soap process.21

    The epoxidation

    of glycidyl ester of soybean oil (GES) was carried out using MCPBA or in situ generated performic acid. The

    former was more efficient, however, due to the low solubility of MCPBA in methylene chloride, large amounts

    of recoverable solvent was required for the epoxidation.

    Figure 1. IR spectra of soybean oil, mixed-FFA, GES and EGS

    Figure 1 shows the FT-IR spectra of soybean oil, mixed FFA, GES and EGS. The band at 3008 cm-1

    was

    attributed to the C-H stretching of =CH in unsaturated fatty acid, such as oleic acid, linoleic acid or linolenic acid.

    New bands at 910 cm-1 and 852 cm-1in the spectrum of GES, with the disappearance at 937 cm-1 in spectrum of

    mixed-FFA means the occurrence of glycidyl group. The conversion of double bonds to epoxy was confirmed by

    the disappearance of the 3008 cm-1 band in GES, and the appearance of band at 752 cm-1 in EGS.

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    7/15

    7

    Figure 2. 1H NMR spectrum and structural assignments of FFA, GE and EGS.

    Table 1. General physical properties of epoxy resins

    Epoxy ResinOxirane oxygen

    (g/100g sample)EEW

    Viscosity at 25oC

    (mPaS)

    EGS 10.1 158 70

    ESO 6.9 232 430

    DGEBA 8.6 186 13000

    Figure 2 shows the1H

    NMR spectra of mixed-FFA, GES and EGS, where linoleic acid was used as a

    generalized compound for structural assignments. The spectra showed no evidence of side reactions in

    preparing GES using soap process, nearly quantitative conversion of double bonds to epoxy groups, and no

    oxirane ring opening during the epoxidation of GES to EGS using MCPBA, i.e., showed complete conversion but

    a lack of side reactions.

    General properties of EGS product compared to ESO, ELO, EGL and DGEBA is shown in Table 1.

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    8/15

    8

    3.2. Viscosity reducing ability

    Figure 3. Viscosity of DGEBA with various EGS/ESO concentrations.

    It was found that EGS had inherently lower viscosity than ESO. EGS has an extra glycidyl group and lower

    molecular weight compared to ESO, which is a triglyceride and has oligomeric behavior. The viscosity reducing

    abilities of EGS and ESO were tested at different concentrations replacement of DGEBA resin, which has a high

    viscosity at 13000 mPaS (see Figure 3). ESO and EGS all showed good miscibility with DGEBA; however, EGS

    exhibited a much better viscosity reducing ability than ESO. Only 30 wt% of EGS reduced the DGEBA resin

    viscosity to value below 1000 mPaS, which is indispensable for many applications. At least 50 wt% of ESO was

    needed to reduce DGEBA resin to the same viscosity.

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    9/15

    9

    3.3. Curing of Reaction

    Figure 4. Dynamic thermograms of DGEBA-EGS/ESO-MHHPA systems

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    10/15

    10

    Table 2. DSC results of curing DGEBA-EGS/ESO-MHHPA systems

    Composition H, (J/g)*H, (kJ/mol)

    Tonset, (

    oC) Tpeak,(

    oC)

    Pure DGEBA 355.3 66.4 133.7 151.910wt% EGS 347.7 63.8 138.0 156.6

    30wt% EGS 348.2 61.8 138.1 157.7

    50wt% EGS 340.2 58.4 136.9 159.2

    70wt% EGS 336.6 55.9 136.5 160.9

    90wt% EGS 334.8 53.9 136.7 163.3

    Pure EGS 321.5 52.5 141.3 167.6

    10wt% ESO 359.9 68.5 134.4 153.9

    30wt% ESO 320.6 63.4 135.1 156.0

    50wt% ESO 300.2 61.7 140.3 159.8

    70wt% ESO 281.4 60.3 143.2 167.8

    90wt% ESO 250.1 55.9 147.6 205.4

    Pure ESO 230.0 52.6 182.6 215.9

    * based on the total number of epoxy groups.

    Differential scanning calorimetry was applied to study the curing behavior of the blended epoxy resins, as

    shown in Figure 3, the exothermic peaks were characteristic of the epoxy and anhydride curing reactions.

    Integration of these peaks allows the determination of the enthalpy of curing reaction(H), onset curing

    temperature (Tonset) and exothermic peaks (Tpeak). The results are shown in Table 2.

    From Figure 3, the pure DGEBA and ESO all show single exothermic reaction peaks at 152oC and 216

    oC,

    respectively. The higher Tpeakvalue of ESO means a slower reaction rate, which was also confirmed by a lower

    H value in Table 2. A lower oxirane content of ESO and also internal epoxy function groups react sluggishly

    with MHHPA curing agent.

    The addition of ESO to DGEBA lead to a shifting ofTpeakand Tonset to higher values. With a decrease ofH

    value, two partially overlapped peaks were clearly observed, especially for 50 wt% ESO or higher replacement,

    which suggested that there are decreasing levels of ESO miscibility in the DGEBA. Non-homogenous mixing will

    prevent complete cure of the epoxy resin. ESO has internal, hindered epoxy groups whereas DGEBA has

    glycidyl groups of less steric hindrance and more reactive than the internal epoxy groups of ESO. Similar results

    have also been reported by Altuna23

    and Boquillon.24

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    11/15

    11

    The prepared EGS resin showed a quite different and interesting curing behavior. The neat EGS shows two

    overlapped peaks, similar to the blend of DGEBA and ESO, which is believed to be due to the inherent different

    reactivity of glycidyl and internal epoxy function groups. The Tpeakand Tonsetvalue of EGS were more than 40oC

    lower than ESO, which indicated EGS is much more reactive than ESO. Increased addition of EGS to DGEBA also

    lead to shifting ofTpeak to higher values, but the Tonset remained nearly constant, and only a 16oC increase of

    Tpeak was observed for 90 wt% EGS replacement compared to pure DGEBA, while it was 54oC for 90 wt% ESO

    replacement.

    The H (J/g) also followed a similar trend. The higher oxirane content of EGS, which bears glycidyl groups

    like DGEBA, appears to facilitate a more homogenous three dimension polymer structure upon curing

    compared to ESO blends. One note of interest was that a small amount of replacement, e.g., 30 wt% EGS or

    below, or 10 wt% ESO, had little effect on the H or Tpeak values of DGEBA, which may be related to

    homogeneity and compatibility with the DGEBA.

    3.4. Thermal Properties

    Figure 5. The change of glass transition temperatures with various ESO/EGS contents

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    12/15

    12

    The glass transition temperature (Tg) is considered one of the fundamental characteristics as it relates to

    polymer properties and processing. For a polymer to serve as a useful plastic, its Tg must be appropriately

    higher than the temperature of its intended work environment.25

    The aliphatic amine,7,26

    or boron trifluoride

    diethyl etherate27,28

    cured ESO polymers generally have very low Tg (even below 0oC). Aromatic amine

    29,

    cycloaliphatic amine,2thermal latent initiator30 or anhydride31 cured ESO polymers generally have higher Tg, but

    it is still rare to see a cured ESO6with Tg above 60

    oC. Low Tg represents a low crosslink density, as mentioned

    above. Due to the low oxirane content of ESO, sluggish reactivity of internal epoxy groups with nucleophilic

    curing agents, the self intermolecular crosslinking of ESO, and unreactive saturated components like stearic

    acid, palmitic acid or myristic acid that act as plasticizers, which further decrease the polymer Tg.

    DSC and Dynamic Mechanical Analysis (DMA) are widely used to characterize the Tg. It is necessary to note

    here that for most thermosetting plastics, the DMA measurement based on the tan peak at a frequency of 1

    Hz generally occurs at a temperature as much as 15-20

    o

    C above Tg as measured by dilatometry or DSC.

    32

    Thechange of cured epoxy resin Tg as measured by DSC is shown in Figure 5.

    Not surprisingly, the MHHPA cured pure EGS had a much higher Tgat 88oC, which was nearly 40

    oC higher

    than ESO though still lower than the pure DGEBA resin.

    Addition of ESO or EGS lead to a decrease ofTg,for smaller replacement, e.g., below 30 wt%, the Tg values

    of ESO-DGEBA or EGS-DGEBA systems were quite similar, which indicated the Tg behavior was mainly

    determined by the DGEBA structure. Further increase the replacement contents of ESO or EGS, theTg values

    decreased rapidly, especially for the ESO system. The inherent aliphatic long chain structure of ESO and the

    higher saturated content and lower epoxy groups make it difficult to produce a densely crosslinked structured

    polymer as some segments will vibrate more freely upon thermal stress.12

    It was also found that neat ESO or higher replacement (above 50 wt%) of ESO-DGEBA thermosetting

    polymer showed a broad transition from the glassy to the rubbery state. Similar behavior was also found in

    ELO replacement of di-glycidyl ether of bisphenol F (DGEBF) resin,33

    which was not found in EGS-DGEBA system.

    The plasticizing effect of saturated fatty acids in the network32and/or the different reactivity of ESO and DGEBA

    lead a broad distribution of polymer molecular weight and indicat a heterogeneous polymer network was

    formed.34

    Figure 6 presents the TGA curves as a function of temperature for the cured epoxy resin. Since ESO-

    DGEBA resin had similar thermal stability with EGS-DGEBA resin, only the latter is shown here. TGA results

    indicated all cured EGS-DGEBA resins appear thermally stable to temperatures below 300 oC. Addition of EGS

    lead to an earlier onset of degradation. All the resins similarly presented two stage degradation behavior. The

    first stage of decomposition, from 350 to 450oC, is believed to be due to the pyrolysis of the crosslinked epoxy

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    13/15

    13

    resin network, decomposition of unreacted MHHPA, and dehydration of hydroxyl groups. The second loss

    stage from ~450 to 600 oC was considered to be the complete decomposition of the smaller fragments like

    cyclized or aromatic degradation byproducts as indicated by a decrease of char residue when EGS component

    was increased in DGEBA.

    Figure 6. TGA analysis of cured EGS-DGEBA blends compared to pure EGS and pure DGEBA.

    3.5. Mechanical PerformanceThe flexural properties of the cured resin system with varying ESO/EGS content were determined and the

    results were listed in Table 3. As can be seen from Table 3, 10 wt% of ESO or EGS replacement of DGEBA lead

    to an improvement in flexural modulus, although further increases of ESO/EGS content lead to a decrease of

    flexural modulus. Similar results were also reported in an amine cured soy-based epoxy resin system.34

    The

    flexural stress of EGS-DGEBA exhibited a gradual decrease until 50 wt%, then followed an abrupt degradation.

    While for ESO-DGEBA, such degradation occurs at 30 wt%.

    As discussed previously, two overlapping peaks in the curing curves determined by DSC were observed

    when high contents of EGS/ESO were added into DGEBA. Different reactivity or reduced compatibility of

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    14/15

    14

    ESO/EGS with DGEBA, may ultimately prevent the ESO/EGS resin from fully participating in the crosslinking.

    Therefore, the resulting crosslinked thermosets may become increasingly plasticized by partially reacted

    ESO/EGS resins at high contents, leading to a decrease of flexural strength. EGS has higher oxirane value and

    was observed to be more reactive with DGEBA compared to ESO, so a higher content of EGS replacement of

    DGEBA was achieved with less degradation of the mechanical strength. Due to the inherent lower epoxy

    content and sluggish reactivity of internal epoxy function groups, a lower ESO content replacement was

    required to prevent additional sacrifice of mechanical performance.

    Table 3. Flexural properties of cured DGEBA-EGS/ESO resin

    Composition Flexural StrengthMPa Flexural ModulusMPa

    Pure DGEBA 138.1 2945.0

    10wt% EGS 133.0 3162.130wt% EGS 125.2 2837.8

    50wt% EGS 121.9 2829.1

    70wt% EGS 91.0 2608.2

    90wt% EGS 81.3 1687.6

    10wt% ESO 124.2 2977.2

    30wt% ESO 120.8 2822.8

    50wt% ESO 107.0 2471.0

    70wt% ESO 81.4 2270.8

    90wt% ESO 60.0 1652.8

    4. ConclusionEGS resin materials were produced from soybean oils with reduced saturated FFA fraction content. The

    products were characterized and showed high oxirane contents and were more reactive than ESO. The EGS

    blends were cured by MHHPA and their thermosetting polymer Tgs measured in comparison to control ESO

    cured in similar fashion. The EGS resin systems had significantly reduced viscosity compared to their pure

    DGEBA or ESO-blend epoxy counterparts. The products displayed glass transitions that were a fairly simple

    function of oxirane content with some added influence of glycidyl versus internal oxirane reactivity. The

    products displayed improved Tgs and mechanical properties compared to their ESO counterparts and, in

    addition to an inherently low viscosity and efficient viscosity reduction, should therefore be more attractive as

  • 7/27/2019 Epoxidized Glycidyl Ester of Soybean Oil as Reactive Diluent For

    15/15

    15

    a reactive diluent. For instance, EGS derived from renewable sources could further enable defect-free

    fabrication of complex, shaped epoxy composites for structural composite applications.

    5. References1 Mark, H. F., Ed. Encyclopedia of Polymer Science and Technology, Wiley-Interscience: New York, 2004.

    2 Czub, P., Macromolecular Symposia242, 60 2006.

    3 Varley, R. J.; Tian, W., Polymer International53, 69 2004.

    4 Raquez, J. M.; Delglise, M.; Lacrampe, M. F.; Krawczak, P., Progress in Polymer Science35, 487 2010.

    5 Cheng, X.; Chen, Y.-X.; Du, Z.-L.; Zhu, P.-X.; Wu, D.-C., Journal of Applied Polymer Science119, 3504 2011.

    6 Tan, S. G.; Chow, W. S., Polymer-Plastics Technology and Engineering49, 1581 2010.

    7 Xua, J. X.; Liu, Z. L.; Erhan, S. Z., Journal of the American Oil Chemists' Society81, 813 2004.

    8 May, C. A., Ed. Epoxy Resins Chemistry and Technology, Marcel Dekker, Inc.: New York, 1988.

    9 Raghavachar, R.; Sarnecki, G.; Baghdachi, J.; Massingill, J., Journal of Coatings Technology72, 125 2000.

    10 Chandrashekhara, K.; Sundararaman, S.; Flanigan, V.; Kapila, S., Materials Science and Engineering: A412, 2 2005.11 Gupta, A. P.; Ahmad, S.; Dev, A., Polymer Engineering & Science51, 1087 2011.

    12 Gelb, L.; Ault, W.; Palm, W.; Witnauer, L.; Port, W., Journal of the American Oil Chemists' Society37, 81 1960.

    13 Czub, P., Macromolecular Symposia245-246, 533 2006.

    14 Kar, S.; Banthia, A. K., Materials and Manufacturing Processes19, 459 2004.

    15 Miyagawa, H.; Misra, M.; Drzal, L. T.; Mohanty, A. K., Polymer Engineering & Science45, 487

    2005.

    16 Jin, F.-L.; Park, S.-J., Materials Science and Engineering: A478, 402 2008.

    17 Tan, S. G.; Chow, W. S., J Therm Anal Calorim101, 1051 2010.

    18 La Scala, J.; Wool, R. P., Polymer46, 61 2005.

    19 Meier, M. A. R.; Metzger, J. O.; Schubert, U. S., Chemical Society Reviews36, 1788 2007.

    20 Gan, L. H.; Ooi, K. S.; Goh, S. H.; Chee, K. K., Journal of Applied Polymer Science46, 329 1992.

    21 Maerker, G.; Saggese, E.; Port, W., Journal of the American Oil Chemists' Society38, 194 1961.

    22 Kester, E. B.; Gaiser, C. J.; Lazar, M. E., The Journal of Organic Chemistry08, 550 1943.

    23 Altuna, F. I.; Espsito, L. H.; Ruseckaite, R. A.; Stefani, P. M., Journal of Applied Polymer Science120, 789 2011.

    24 Boquillon, N.; Fringant, C., Polymer41, 8603 2000.

    25 Stevens, M. P., Polymer Chemistry : An Introduction; Oxford University Press 1999.

    26 Lu, P.; University of Missouri-Rolla, Rolla: 2001.

    27 Liu, Z.; Doll, K. M.; Holser, R. A., Green Chemistry11, 1774 2009.

    28 Liu, Z.; Erhan, S. Z., Journal of the American Oil Chemists' Society87, 437 2009.

    29 Earls, J. D.; White, J. E.; Lpez, L. C.; Lysenko, Z.; Dettloff, M. L.; Null, M. J., Polymer48, 712 2007.

    30 Park, S.-J.; Jin, F.-L.; Lee, J.-R., Macromolecular Rapid Communications25, 724 2004.

    31 Gerbase, A.; Petzhold, C.; Costa, A., Journal of the American Oil Chemists' Society79, 797 2002.

    32 Tanrattanakul, V.; Saithai, P., Journal of Applied Polymer Science114, 3057 2009.

    33 Miyagawa, H.; Mohanty, A. K.; Misra, M.; Drzal, L. T., Macromolecular Materials and Engineering289, 629 2004.

    34 Miyagawa, H.; Misra, M.; Drzal, L. T.; Mohanty, A. K., Journal of Polymers and the Environment13, 87 2005.


Recommended