Next Article in Journal
Green Aspects in Molecularly Imprinted Polymers by Biomass Waste Utilization
Next Article in Special Issue
Analysis of the Electrical and Thermal Properties for Magnetic Fe3O4-Coated SiC-Filled Epoxy Composites
Previous Article in Journal
Conjugated Polymer Films Having a Uniaxial Molecular Orientation and Network Structure Prepared by Electrochemical Polymerization in Liquid Crystals
Previous Article in Special Issue
Probabilistic Assessment of Fracture Toughness of Epoxy Resin EPOLAM 2025 Including the Notch Radii Effect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Reaction Mechanism and Mechanical Property Improvement of Poly(Lactic Acid) Reactive Blending with Epoxy Resin

by
Krittameth Kiattipornpithak
1,
Nanthicha Thajai
2,
Thidarat Kanthiya
1,
Pornchai Rachtanapun
1,3,4,
Noppol Leksawasdi
1,3,4,
Yuthana Phimolsiripol
1,3,4,
David Rohindra
5,
Warintorn Ruksiriwanich
4,6,
Sarana Rose Sommano
4,7 and
Kittisak Jantanasakulwong
1,3,4,*
1
Faculty of Agro-Industry, Chiang Mai University, Mae-Hea, Mueang, Chiang Mai 50100, Thailand
2
Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
3
Cluster of Agro Bio-Circular-Green Industry, Chiang Mai University, Chiang Mai 50100, Thailand
4
Center of Excellence in Materials Science and Technology, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
5
School of Biological and Chemical Sciences, Faculty of Science, Technology and Environment, The University of the South Pacific, Laucala Bay Road, Suva, Fiji
6
Department of Pharmaceutical Sciences, Faculty of Pharmacy, Chiang Mai University, Mae-Hia, Muang, Chiang Mai 50100, Thailand
7
Plant Bioactive Compound Laboratory (BAC), Department of Plant and Soil Sciences, Faculty of Agriculture, Chiang Mai University, Chiang Mai 50200, Thailand
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(15), 2429; https://doi.org/10.3390/polym13152429
Submission received: 3 July 2021 / Revised: 18 July 2021 / Accepted: 18 July 2021 / Published: 23 July 2021
(This article belongs to the Special Issue Advanced Epoxy-Based Materials II)

Abstract

:
Polylactic acid (PLA) was melt-blended with epoxy resin to study the effects of the reaction on the mechanical and thermal properties of the PLA. The addition of 0.5% (wt/wt) epoxy to PLA increased the maximum tensile strength of PLA (57.5 MPa) to 67 MPa, whereas the 20% epoxy improved the elongation at break to 12%, due to crosslinking caused by the epoxy reaction. The morphology of the PLA/epoxy blends showed epoxy nanoparticle dispersion in the PLA matrix that presented a smooth fracture surface with a high epoxy content. The glass transition temperature of PLA decreased with an increasing epoxy content owing to the partial miscibility between PLA and the epoxy resin. The Vicat softening temperature of the PLA was 59 °C and increased to 64.6 °C for 0.5% epoxy. NMR confirmed the reaction between the -COOH groups of PLA and the epoxy groups of the epoxy resin. This reaction, and partial miscibility of the PLA/epoxy blend, improved the interfacial crosslinking, morphology, thermal properties, and mechanical properties of the blends.

1. Introduction

Biodegradable polymers have attracted considerable attention in recent years owing to their environmental effects. Petroleum plastics are being replaced by renewable, eco-friendly materials. Biodegradable polymers, such as PLA [1], polybutylene succinate (PBS) [2], thermoplastic starch (TPS) [3], starch [4], polysaccharides [5], carboxymethyl bacterial cellulose [6], and pectin [7,8] have been widely studied. PLA is an eco-friendly polymer as it is synthesized using lactic acid extracted from natural sources. PLA is a typical biodegradable plastic [9] that can be used to replace commercial single-use plastics [10,11]. PLA is one of most promising candidates to replace petroleum plastic as the monomers are produced from renewable feedstock. PLA is synthesized from lactic acid using ring opening polymerization, polycondensation, and enzymatic polymerization [12]. PLA is a thermoplastic material, with high-mechanical properties, biodegradability, biocompatibility, and transparency. However, PLA has disadvantages such as brittleness, heat resistance, and barrier properties [13]. Property improvement of PLA has been investigated by several researchers [14,15,16]. Several studies have investigated the improvement of PLA brittleness and flexibility [14]. Polymer blending and co-polymerization are two methods that can be used to improve polymer mechanical properties [17,18]. Notably, PLA has been previously reported as a coating material to improve mechanical properties and water resistance of TPS [19].
Epoxy resin is a reactive polymer that contains epoxide groups in the structure, which are used to crosslink with variant functional groups such as amides and carboxylic [20]. Bisphenol A based epoxy resin is the common epoxy resin, which is derived by the reaction of thiol, alcohols, and amines with epichlorohydrin [20]. Epoxy resin is a substance that is used to improve PLA owing to its increased toughness, low shrinkage, corrosion resistance, and improved heat resistance [21,22,23,24]. Epoxy resins have epoxy groups that react with amines [25], carboxylic acids, and hydroxyl groups [26]. Epoxy interacts with some thermoplastics, such as polycarbonate [27], polyamine [28], and PLA [29]. Epoxy resin acts as a bridge to connect the polymer on both reactive sides, thus allowing the epoxy to bind different polymers [26]. In addition, chemical reactions influence the morphology, melting, and crystallization behaviors that lead to property changes [10]. Nevertheless, a high amount of epoxy resin creates a highly networked structure, resulting in a thermoset polymer. Incorporation of epoxy resin into PLA increased molecular weight, and networked and branched structures which slowed down the hydrolytic degradation of PLA by approximately 130% compared to pure PLA [30]. Epoxy resin was used as a compatibilizer to improve compatibility of the PLA/polyamide blend [31]. Reactive blending of epoxy resin and PLA increased mechanical properties, melt strength, torque, toughness, and melt viscosity, but decreased crystallinity of PLA [32,33]. These improvements were due to a reaction between -COOH end groups of PLA and epoxy groups of epoxy resin [32,33]. However, the reaction mechanism, plasticizing effect, and morphology of epoxy resin blending with PLA have not been reported.
Therefore, in this study, the effect of epoxy resin on PLA blending was investigated. PLA was melt-blended with 0–20% epoxy at 170 °C for 10 min. Mechanical properties, morphology, thermal properties, thermal stability, and the reaction mechanism were investigated.

2. Materials and Methods

2.1. Materials

Poly(lactic acid) (PLA) pellet (4032D, MW 100,000 g/mol, density 1.24 g/cc, MFI = 7 g/10 min at 210 °C, NatureWorks LLC., Minnetonka, MN, USA) was purchased from PTT Global Chemical Pub Co., Ltd., Bangkok, Thailand. Diglycidyl ether of bisphenol A epoxy resin (grade 0302, liquid state) was purchased from EASY Resin Co., Ltd., Nonthaburi, Thailand.

2.2. Sample Preparation

PLA was blended with epoxy resin using a two-roll mill machine (Model PII140, Pirom‒Olarn, Bangkok, Thailand) at 170 °C for 10 min, and then compressed into sheets by a hot compress at 170 °C for 10 min, followed by quenching at 10–25 °C. PLA was blended with 0.5–20% (wt/wt), a monomer of epoxy resin without a hardener to study the effects of the epoxy content. Code name and composition of the sample are shown in Table 1.

2.3. Tensile Properties

The tensile properties were measured following JISK-6251-7 using a tensile tester at a crosshead speed of 10 mm/min with a gauge length of 10 mm. Bone-shaped specimens of 30 × 10 × 0.2 mm (length × width × thickness) were prepared using compression molding at 170 °C for 5 min.

2.4. Scanning Electron Microscopy (SEM)

Morphologies of the blended samples were characterized using SEM (JSM-5910LV JEOL Co., Ltd., Tokyo, Japan) at 15 kV. The samples were broken in liquid nitrogen, followed by the coating of the fracture surface with a thin layer of gold using sputtering (108 Auto/SE sputter coater, Cressington Co., Ltd., Watford, England).

2.5. Differential Scanning Calorimetry (DSC)

% X c =( Δ H m Δ H c H m 0   )×100
where ΔHm and ΔHc are the enthalpy of melting and cold crystallization, respectively. Δ H m 0 is the melting enthalpy of 100% PLA (93.7 J/g) [34].

2.6. Vicat Softening Temperature

The samples with dimensions of 10 mm × 10 mm × 3 mm (width × length × thickness) were prepared using hot-compression molding at 170 °C for 5 min. The prepared samples were then tested by increasing the temperature until the flattened needle penetrated 1 mm into the surface using the ASTM D1525 standard. At least five specimens of each sample were tested.

2.7. Nuclear Magnetic Resonance (NMR)

The spectra were acquired using NMR (NEOTM 500 MHz, Bruker Co., Ltd., Boston, MA, USA). Samples were dissolved in a chloroform solvent (CDCL3) at 25 mg/mL before observation. Analysis of NMR intensities via different statistical models was evaluated using a custom-written Topspin 4.0.8 (Bruker BioSpin GmbH, Karlsruhe, Germany).

2.8. Statistical Analysis

The one-way ANOVA with the Statistical Package for the Social Sciences, SPSS Version 17 (SPSS, Armonk, NY, USA) was used to analyze the data. The differences found (p < 0.05) were evaluated using Duncan’s test.

3. Results and Discussion

3.1. Mechanical Properties

PLA was melt-blended with epoxy resin (0–20% w/w) to observe the effect of epoxy on the mechanical properties of the blends. The stress–strain curve, maximum tensile strength, and elongation at break are shown in Figure 1a,b. The maximum tensile strength of the neat PLA was 57.5 MPa, and the addition of 0.5% epoxy enhanced the maximum tensile strength to 66.9 MPa due to crosslinking caused by the epoxy reaction. Increasing the epoxy content extended the elongation at break of the blends, as a high crosslinking inside the PLA phase led to the formation of a network structure [35]. The Young’s modulus of PLA was 2.3 GPa, while PLA blend with epoxy 0.5, 1, 2, 5, 10, and 20% were 1.6, 1.7, 1.5, 1.6, 1.2, and 1.5 GPa, respectively. The epoxy resin reduced Young’s modulus of PLA, owing to the partial miscibility between PLA and the epoxy [24]. The high amount of epoxy acted as a plasticizer for the PLA, thereby reducing Young’s modulus and maximum tensile strength, and increasing the elongation at break, owing to its effect on the enhanced distance between the PLA molecules [29].

3.2. Morphology

Figure 2 shows the morphology of the PLA blend with 0.5–20% epoxy resin. The PLA/epoxy0.5 blend exhibited small epoxy particles (~200 nm) dispersed in the PLA matrix. Holes of the removed epoxy particles from the PLA matrix were also observed (Figure 2a). The formation of epoxy nanoparticles indicated a high compatibility between PLA and the epoxy. The PLA blend with 1–2% epoxy exhibited fine epoxy nanoparticles distributed in the PLA matrix without the removal of the epoxy particles, whereas a high epoxy content (5–20%) suggested nanoparticles that were smaller than those in the PLA/epoxy0.5 blend with a smooth fracture surface. This can be attributed to the high interfacial adhesion between PLA and the epoxy through the crosslinking interfacial reaction between PLA and the epoxy [24]. The small sizes of the epoxy particles and high interfacial adhesion resulted in the high transparency of the blend due to low light scattering [36].

3.3. Differential Scanning Calorimetry (DSC)

DSC curves were used to determine the effect of the epoxy resin on the thermal properties of the PLA/epoxy blends. The Tg and Tm were measured during the second heating scan. Figure 3 shows the DSC curves of the PLA, epoxy, and PLA blend with 0.5–20% epoxy. The Tg and Tm of PLA were 60 °C and 166 °C and tended to decrease with increasing epoxy content owing to small crystal sizes [37] and partial miscibility between the epoxy and PLA blends, respectively [38]. The decrease in Tg indicated that the epoxy acts as a plasticizer for the PLA. Large exothermic peak of the PLA/epoxy blends indicated to recrystallization during second heating scan of DSC measurement, which presented large endothermic peak of recrystallization at 110–135 °C. This recrystallization was not observed in pure PLA.
The addition of the epoxy led to an increase in the chain length and reduced the mobility of the PLA chain [39]. The PLA crosslinked structure prevented the formation of inter- and intramolecular interactions of PLA crystallinity [40]; this resulted in the low crystallinity (0.6–3.6%) of the blends. The epoxy reaction reduced the number of PLA chain terminals in the structure and prevented the formation of nuclei as well as the growth of the crystals [29,41].

3.4. Vicat Softening Temperature (VST)

The VST test was used to determine the thermal stability based on the heat distortion temperature. The VST of PLA and the PLA blend with 0.5–20% epoxy resin are depicted in Figure 4. The VST of the neat PLA was 59 °C, and when 0.5–2% epoxy was added, the VST showed an increasing trend to 64.6 °C. The VST of the PLA/epoxy20 was reduced to 55.7 °C. The increased VST of the PLA/epoxy2 blend was due to the network structure of PLA, that was a result of the crosslinking reaction. Improvement of the VST due to internal crosslinking has been previously reported [42]. The decreased VST in the PLA blend with 5–20% epoxy indicated an excessive amount of epoxy and the crosslinking density inside PLA that reduced the crystal formation of the PLA. Thermal stability improvement due to a crosslinking structure and crystal formation has been previously reported [43].

3.5. Reaction Mechanism

Figure 5 shows the 13C NMR spectra of the PLA, epoxy, and the PLA/epoxy20 blend. The epoxy showed peaks of -CH3 (bisphenol A) at 30.7, oxirane ring carbons at 43.9 and 49.7 ppm, and bisphenol A carbons at 41.2, 68.0, 113.6, 127.3, 143.1, and 156.0 ppm [44]. Neat PLA showed peaks at 16.7 (-CH3), 69.1 (methylene carbon), and 169.70 ppm (-C=O) [45]. The 13C NMR spectra of the PLA/epoxy blends showed the characteristic peaks of PLA and epoxy resin at the same position as that of neat epoxy and PLA. New peaks were observed at 21 and 67 ppm, corresponding to C10 and C11, respectively. Figure 6 shows the 1H NMR spectra of PLA, epoxy, and the PLA/epoxy20 blend. Epoxy showed peaks -CH3 of bisphenol A at 1.6 ppm (b), -CH2 of oxirane ring at 2.6 and 2.8 ppm (c), -CH of oxirane ring at 3.29 ppm (d), -CH2 at 3.9 and 4.29 ppm (e), and aromatic protons of bisphenol A at 6.8 and 7.1 ppm [44]. Neat PLA showed peaks at 1.6 (a,-CH3) and 5.1 ppm (b,-CH). The 1H NMR spectra of the PLA/epoxy blend showed characteristic peaks of -CH3 (1.6) and -CH (5.1) at the same position as that of neat PLA. Furthermore, the 1H epoxy peaks of bisphenol A (1.6, 6.8, and 7.1) shifted to lower positions, while the peaks of the oxirane ring (2.6, 2.8, 3.29, 3.9, and 4.29) shifted to higher positions. New peaks were observed at 2.57 and 3.40 which indicated H7 and H6, respectively. The shifting of bisphenol A and the oxirane ring, and the appearance of two new peaks, confirmed the reaction between PLA and the epoxy. This also suggested a reaction between the epoxy groups of the epoxy resin and the -COOH end groups of PLA (Figure 6c). Previous studies have reported on the reaction between epoxy groups and -COOH groups [46,47]. This reaction improved the tensile properties, toughness, morphology, and thermal properties of PLA.

4. Conclusions

Epoxy resin was successfully blended with PLA to improve the mechanical properties and reaction mechanism of the blends. The maximum tensile strength of PLA was improved from 57.5 MPa (neat PLA) to 66.9 MPa with 0.5% epoxy, whereas the elongation at break showed a significant increase with 20% epoxy. Morphology of the PLA/epoxy blends showed nanoparticles dispersed in the PLA matrix, while a smooth fracture surface of the 5–10% PLA/epoxy blends was observed due to the high interfacial adhesion between PLA and epoxy. The Tm and Tg of the PLA/epoxy blends decreased with increasing epoxy content, owing to the nucleating effect of small epoxy particle sizes, and the partial miscibility between PLA and epoxy acting as a plasticizer, respectively. The VST of PLA increased with 2% epoxy, while an excessive amount of epoxy reduced the VST due to reduced crystal formation. NMR results confirmed the reaction between the -COOH groups of PLA and the epoxy groups of the epoxy resin. This reaction improved the mechanical properties, toughness, morphology, and thermal properties of the blends, additionally resulting in high optical transparency. The PLA/epoxy blends also contained unreacted epoxy groups that could react with other reactive functional groups for reactive blending as a compatibilizer. PLA/epoxy can be applied for packaging, medical, and agriculture applications.

Author Contributions

Investigation, data curation and writing‒original draft preparation, K.K.; investigation, N.T. and T.K.; visualization and writing‒review and editing, P.R., N.L., Y.P., D.R., W.R. and S.R.S.; Conceptualization, writing‒original draft preparation, project administration, funding acquisition and writing‒review and editing, K.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Program Management Unit for Human Resources & Institutional Development, Research and Innovation, Office of National Higher Education Science Research and Innovation Policy Council (NXPO) [Grant Number B16F640001]. This research was partially supported by the Chiang Mai University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors gratefully acknowledge the Faculty of Agro-Industry, Chiang Mai University, for their support.

Conflicts of Interest

The authors declare no competing interests.

Abbreviations

PLApolylactic acid
PBSpolybutylene succinate
TPSthermoplastic starch
SEMscanning electron microscopy
DSCdifferential scanning calorimetry
Tmmelting temperature
Tgglass transition temperature
NMRnuclear magnetic resonance
CDCL3chloroform solvent

References

  1. Gregor-Svetec, D.; Leskovšek, M.; Leskovar, B.; Elesini, U.S.; Leskovar, B.; Elesini, U.S.; Vrabič-Brodnjak, U. Analysis of PLA composite filaments reinforced with lignin and polymerised-lignin-treated NFC. Polymers 2021, 13, 2174. [Google Scholar] [CrossRef]
  2. Jantanasakulwong, K.; Rohindra, D.; Mori, K.; Kuboyama, K.; Ougizawa, T. Thermoplastic elastomer by reactive blending of poly(butylene succinate) with ethylene-propylene-diene terpolymer and ethylene-1-butene rubbers. J. Elastomers Plast. 2015, 47, 215–231. [Google Scholar] [CrossRef]
  3. Jumaidin, R.; Diah, N.A.; IIyas, R.A.; Alamjuri, R.H.; Yusof, F.A.M. Processing and characterization of banana leaf fibre reinforced thermoplastic cassava starch composites. Polymers 2021, 13, 1420. [Google Scholar] [CrossRef] [PubMed]
  4. Malyszek, Z.; Lewandowicz, J.; Thanh-Blicharz, J.L.; Walkowiak, K.; Kowalczewski, P.L.; Baranowska, H.M. Water behavior of emulsions stabilized by modified potato starch. Polymers 2021, 13, 2200. [Google Scholar] [CrossRef] [PubMed]
  5. Surin, S.; You, S.; Seesuriyachan, P.; Muangrat, R.; Wangtueai, S.; Jambrak, A.R.; Phongthai, S.; Jantanasakulwong, K.; Chaiyaso, T.; Phimolsiripol, Y. Optimization of ultrasonic-assisted extraction of polysaccharides from purple glutinous rice bran (Oryza sativa L.) and their antioxidant activities. Sci. Rep. 2020, 10, 10410. [Google Scholar] [CrossRef] [PubMed]
  6. Rachtanapun, P.; Klunklin, W.; Jantrawut, P.; Leksawasdi, N.; Jantanasakulwong, K.; Phimolsiripol, Y.; Seesuriyachan, P.; Chaiyaso, T.; Ruksiriwanich, W.; Phongthai, S.; et al. Effect of monochloroacetic acid on properties of carboxymethyl bacterial cellulose powder and film from nata de coco. Polymers 2021, 13, 488. [Google Scholar] [CrossRef] [PubMed]
  7. Doan, C.T.; Chen, C.L.; Nguyen, V.B.; Tran, T.N.; Nguyen, A.D.; Wang, S.L. Conversion of pectin-containing by-products to pectinases by Bacillus amyloliquefaciens and its applications on hydrolyzing banana pells for prebiotics production. Polymers 2021, 13, 1483. [Google Scholar] [CrossRef] [PubMed]
  8. Wongkaew, M.; Sommano, S.R.; Tangpao, T.; Rachtanapun, P.; Jantanasakulwong, K. Mango peel pectin by microwave-assisted extraction and its use as fat replacement in dried Chinese sausage. Foods 2020, 9, 450. [Google Scholar] [CrossRef] [Green Version]
  9. Qi, X.; Ren, Y.; Wang, X. New advances in the biodegradation of Poly(lactic) acid. Int. Biodeterior. Biodegrad. 2017, 117, 215–223. [Google Scholar] [CrossRef]
  10. Thomas, R.; Yumei, D.; Yuelong, H.; Le, Y.; Moldenaers, P.; Weimin, Y.; Czigany, T.; Thomas, S. Miscibility, morphology, thermal, and mechanical properties of a DGEBA based epoxy resin toughened with a liquid rubber. Polymer 2008, 49, 278–294. [Google Scholar] [CrossRef]
  11. Lee, J.; McCarthy, S. Biodegradable poly(lactic acid) blends with chemically modified polyhydroxyoctanoate through chain extension. J. Polym. Environ. 2009, 17, 240. [Google Scholar] [CrossRef]
  12. Garlotta, D. A literature review of poly(lactic acid). J. Polym. Environ. 2001, 9, 63–84. [Google Scholar] [CrossRef]
  13. Farah, S.; Anderson, D.G.; Langer, R. Physical and mechanical properties of PLA, and their functions in widespread applications—Comprehensive review. Adv. Drug Deliv. Rev. 2016, 107, 367–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Jantanasakulwong, K.; Kobayashi, Y.; Kuboyama, K.; Ougizawa, T. Thermoplastic vulcanizate based on poly(lactic acid) and acrylic rubber blended with ethylene ionomer. J. Macromol. Sci. Phys. Part B 2016, 55, 1068–1085. [Google Scholar] [CrossRef]
  15. Tokiwa, Y.; Calabia, B.P. Biodegradability and biodegradation of poly(lactide). Appl. Microbiol. Biotechnol. 2006, 72, 244–251. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, X.; Hu, H.; Wang, X.; Yu, X.; Zhou, W.; Peng, S. Super tough poly(lactic acid) blends: A comprehensive review. RSC Adv. 2020, 10, 13316–13368. [Google Scholar] [CrossRef] [Green Version]
  17. Su, S.; Kopitzky, R.; Tolga, S.; Kabasci, S. Polylactide (PLA) and its blends with poly(butylene succinate) (PBS): A brief review. Polymers 2019, 11, 1193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Chavalitpanya, K.; Phattanarudee, S. Poly(lactic acid)/Polycaprolactone Blends Compatibilized with Block Copolymer. Energy Procedia 2013, 34, 542–548. [Google Scholar] [CrossRef] [Green Version]
  19. Jantanasakulwong, K.; Homsaard, N.; Phengchan, P.; Rachtanapun, P.; Leksawasdi, N.; Phimolsiripol, Y.; Techapun, C.; Jantrawut, P. Effect of dip coating polymer solutions on properties of thermoplastic cassava starch. Polymers 2019, 11, 1746. [Google Scholar] [CrossRef] [Green Version]
  20. Verma, C.; Olasunkanmi, L.O.; Akpan, E.D.; Quraishi, M.A.; Dagdag, O.; Gouri, M.E.; Sherif, E.-S.M.; Ebenso, E.E. Epoxy resins as anticorrosive polymeric materials: A review. React. Funct. Polym. 2020, 156, 104741. [Google Scholar] [CrossRef]
  21. Bucknall, C.B.; Gilbert, A.H. Toughening tetrafunctional epoxy resins using polyetherimide. Polymer 1989, 30, 213–217. [Google Scholar] [CrossRef]
  22. Hourston, D.J.; Lane, J.M. The toughening of epoxy resins with thermoplastics: 1. Trifunctional epoxy resin-polyetherimide blends. Polymer 1992, 33, 1379–1383. [Google Scholar] [CrossRef]
  23. Park, S.J.; Kim, H.C. Thermal stability and toughening of epoxy resin with polysulfone resin. J. Polym. Sci. Part B Polym. Phys. 2001, 39, 121–128. [Google Scholar] [CrossRef]
  24. Silva, I.D.S.; Barros, J.J.P.; Albuquerque, A.; Jaques, N.G.; Fook, M.V.L.; Wellen, R.M.R. Insights into the curing kinetics of epoxy/PLA: Implications of the networking structure. Express Polym. Lett. 2020, 14, 1180–1196. [Google Scholar] [CrossRef]
  25. Kodsangma, A.; Homsaard, N.; Nadon, S.; Rachtanapun, P.; Leksawasdi, N.; Phimolsiripol, Y.; Insomphun, C.; Seesuriyachan, P.; Chaiyaso, T.; Jantrawut, P.; et al. Effect of sodium benzoate and chlorhexidine gluconate on a biothermoplastic elastomer made from thermoplastic starch-chitosan blended with epoxidized natural rubber. Carbohydr. Polym. 2020, 242, 116421. [Google Scholar] [CrossRef]
  26. Hu, L.; Vuillaume, P.Y. Chapter 7—Reactive compatibilization of polymer blends by coupling agents and interchange catalysts. In Compatibilization of Polymer Blends, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 205–248. [Google Scholar]
  27. Choi, J.; Lim, S.; Kim, J.; Choe, C.R. Studies of an epoxy-compatibilized poly(phenylene sulfide)/polycarbonate blend. Polymer 1997, 38, 4401–4406. [Google Scholar] [CrossRef]
  28. Chiou, K.C.; Chang, F.C. Reactive compatibilization of polyamide-6 (PA 6)/polybutylene terephthalate (PBT) blends by a multifunctional epoxy resin. J. Polym. Sci. Part B Polym. Phys. 2000, 38, 23–33. [Google Scholar] [CrossRef]
  29. Liu, Z.-Y.; Weng, Y.-X.; Huang, Z.-G.; Wang, L.; Qiu, D.; Shao, S.-X. Effect of epoxy chain extender on the properties of polylactic acid. J. Appl. Mater. Sci. Eng. Res. 2019, 2, 1–7. [Google Scholar]
  30. Limsukon, W.; Auras, R.; Selke, S. Hydrolytic degradation and lifetime prediction of poly(lactic acid) modified with a multifunctional epoxy-based chain extender. Polym. Test. 2019, 80, 106108. [Google Scholar] [CrossRef]
  31. Yu, X.; Wang, X.; Zhang, Z.; Peng, S.; Chen, H.; Zhao, X. High-performance fully bio-based poly(lactic acid)/polyamide11 (PLA/PA11) blends by reactive blending with multi functionalized epoxy. Polym. Test. 2019, 78, 105980. [Google Scholar] [CrossRef]
  32. Corre, Y.-M.; Duchet, J.; Reignier, J.; Maazouz, A. Melt strengthening of poly (lactic acid) through reactive extrusion with epoxy-functionalized chains. Rheol. Acta 2011, 50, 613–629. [Google Scholar] [CrossRef]
  33. Karkhanis, S.S.; Matuana, L.M. Extrusion blown films of poly(lactic acid) chain-extended with food grade multifunctional epoxies. Polym. Eng. Sci. 2019, 59, 2211–2219. [Google Scholar] [CrossRef]
  34. Jia, S.; Yu, D.; Zhu, Y.; Wang, Z.; Chen, L.; Fu, L. Morphology, crystallization and thermal behaviors of PLA-based composite: Wonderful effects of hybrid GO/PEG via Dynamic impregnating. Polymers 2017, 9, 528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Semba, T.; Kitagawa, K.; Ishiaku, U.S.; Hamada, H. The Effect of Crosslinking on the Mechanical Properties of Polylactic Acid/Polycaprolactone Blends. J. Appl. Polym. Sci. 2006, 101, 1816–1825. [Google Scholar] [CrossRef]
  36. Maruhashi, Y.; Iida, S. Transparency of polymer blends. Polym. Eng. Sci. 2001, 41, 1987–1995. [Google Scholar] [CrossRef]
  37. Aumnate, C.; Rudolph, N.; Sarmadi, M. Recycling of polypropylene/polyethylene blends: Effect of chain structure on the crystallization behaviors. Polymers 2019, 11, 1456. [Google Scholar] [CrossRef] [Green Version]
  38. Aid, S.; Eddhahak, A.; Ortega, Z.; Froelich, D.; Tcharkhtchi, A. Experimental study of the miscibility of ABS/PC polymer blends and investigation of the processing effect. J. Appl. Polym. Sci. 2017, 134, 44975. [Google Scholar] [CrossRef]
  39. Bednarek, M.; Borska, K.; Kubisa, P. Crosslinking of polylactide by high energy irradiation and photo-curing. Molecules 2020, 25, 4919. [Google Scholar] [CrossRef] [PubMed]
  40. Simis, K.S.; Bistolfi, A.; Bellare, A.; Pruitt, L.A. The combined effects of crosslinking and high crystallinity on the microstructural and mechanical properties of ultra high molecular weight polyethylene. Biomaterials 2006, 27, 1688–1694. [Google Scholar] [CrossRef]
  41. Zamponi, M.; Monkenbusch, M.; Willner, L.; Wischnewski, A.; Farago, B.; Richter, D. Contour length fluctuations in polymer melts: A direct molecular proof. Europhys. Lett. 2005, 72, 1039–1044. [Google Scholar] [CrossRef]
  42. Garcıa-Castaneda, C.; Benavides, R.; Martinez-Pardo, M.E.; Uribe, R.M.; Carrasco-Abrego, H.; Martinez, G. Crosslinking of rigid PVC by ionizing radiation to improve its thermal properties. Radiat. Phys. Chem. 2010, 79, 335–338. [Google Scholar] [CrossRef]
  43. Xiang, H.X.; Zabihi, F.; Zhang, X.Z.; Zhu, M.F. The Crystallization, Melting Behaviors and Thermal Stability of Cross-linked Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Octavinyloctasilasesquioxane. Chin. J. Polym. Sci. 2018, 36, 1353–1360. [Google Scholar] [CrossRef]
  44. Ahmad, S.; Gupta, A.P.; Sharmin, E.; Alam, M.; Pandey, S.K. Synthesis, characterization and development of high performance siloxane-modified epoxy paints. Prog. Org. Coat. 2005, 54, 248–255. [Google Scholar] [CrossRef]
  45. Klinkajorn, J.; Tanrattanakul, V. The effect of epoxide content on compatibility of poly(lactic acid)/epoxidized natural rubber blends. J. Appl. Polym. Sci. 2020, 137, 48996. [Google Scholar] [CrossRef]
  46. Blank, W.J.; He, Z.A.; Picci, M. Catalysis of the epoxy-carboxyl reaction. J. Coat. Technol. 2002, 74, 33–41. [Google Scholar] [CrossRef]
  47. Morancho, J.M.; Ramis, X.; Fernandez-Francos, X.; Konuray, O.; Salla, J.M.; Serra, A. Dual curing of an epoxy resin with dicarboxylic acids. J. Therm. Anal. Calorim. 2020, 142, 607–615. [Google Scholar] [CrossRef]
Figure 1. Tensile properties of PLA blend with 0–20% epoxy (a) Stress–strain curve, and (b) Maximum tensile strength and elongation at break. The different of each lowercase and each uppercase letters indicate the mean values of the elongation at break and the maximum tensile strength differ significantly (p < 0.05), respectively.
Figure 1. Tensile properties of PLA blend with 0–20% epoxy (a) Stress–strain curve, and (b) Maximum tensile strength and elongation at break. The different of each lowercase and each uppercase letters indicate the mean values of the elongation at break and the maximum tensile strength differ significantly (p < 0.05), respectively.
Polymers 13 02429 g001
Figure 2. SEM fracture surface images of (a) PLA/epoxy0.5, (b) PLA/epoxy1, (c) PLA/Epoxy2, (d) PLA/epoxy5, (e) PLA/Epoxy10, and (f) PLA/Epoxy20.
Figure 2. SEM fracture surface images of (a) PLA/epoxy0.5, (b) PLA/epoxy1, (c) PLA/Epoxy2, (d) PLA/epoxy5, (e) PLA/Epoxy10, and (f) PLA/Epoxy20.
Polymers 13 02429 g002
Figure 3. The second scan DSC curves of epoxy, PLA, and the PLA blend with 0.5–20% epoxy.
Figure 3. The second scan DSC curves of epoxy, PLA, and the PLA blend with 0.5–20% epoxy.
Polymers 13 02429 g003
Figure 4. The Vicat softening temperature of PLA and the PLA blend with 0.5–20% epoxy. The mean values indicate by the different lowercase letters differ significantly (p < 0.05).
Figure 4. The Vicat softening temperature of PLA and the PLA blend with 0.5–20% epoxy. The mean values indicate by the different lowercase letters differ significantly (p < 0.05).
Polymers 13 02429 g004
Figure 5. 13C NMR spectra and structures of (a) neat PLA, (b) epoxy, and (c) the PLA/epoxy20 blend.
Figure 5. 13C NMR spectra and structures of (a) neat PLA, (b) epoxy, and (c) the PLA/epoxy20 blend.
Polymers 13 02429 g005
Figure 6. 1H NMR spectra and structures of (a) neat PLA, (b) epoxy, and (c) the PLA/epoxy20 blend.
Figure 6. 1H NMR spectra and structures of (a) neat PLA, (b) epoxy, and (c) the PLA/epoxy20 blend.
Polymers 13 02429 g006
Table 1. Code name and composition of PLA and epoxy resin blend.
Table 1. Code name and composition of PLA and epoxy resin blend.
SamplePLAEpoxy Resin
PLA1000
PLA/epoxy0.599.50.5
PLA/epoxy1991
PLA/epoxy2982
PLA/epoxy5955
PLA/epoxy109010
PLA/epoxy208020
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kiattipornpithak, K.; Thajai, N.; Kanthiya, T.; Rachtanapun, P.; Leksawasdi, N.; Phimolsiripol, Y.; Rohindra, D.; Ruksiriwanich, W.; Sommano, S.R.; Jantanasakulwong, K. Reaction Mechanism and Mechanical Property Improvement of Poly(Lactic Acid) Reactive Blending with Epoxy Resin. Polymers 2021, 13, 2429. https://doi.org/10.3390/polym13152429

AMA Style

Kiattipornpithak K, Thajai N, Kanthiya T, Rachtanapun P, Leksawasdi N, Phimolsiripol Y, Rohindra D, Ruksiriwanich W, Sommano SR, Jantanasakulwong K. Reaction Mechanism and Mechanical Property Improvement of Poly(Lactic Acid) Reactive Blending with Epoxy Resin. Polymers. 2021; 13(15):2429. https://doi.org/10.3390/polym13152429

Chicago/Turabian Style

Kiattipornpithak, Krittameth, Nanthicha Thajai, Thidarat Kanthiya, Pornchai Rachtanapun, Noppol Leksawasdi, Yuthana Phimolsiripol, David Rohindra, Warintorn Ruksiriwanich, Sarana Rose Sommano, and Kittisak Jantanasakulwong. 2021. "Reaction Mechanism and Mechanical Property Improvement of Poly(Lactic Acid) Reactive Blending with Epoxy Resin" Polymers 13, no. 15: 2429. https://doi.org/10.3390/polym13152429

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop