Next Article in Journal
Re-Assemblable, Recyclable, and Self-Healing Epoxy Resin Adhesive Based on Dynamic Boronic Esters
Next Article in Special Issue
Forefront Research of Foaming Strategies on Biodegradable Polymers and Their Composites by Thermal or Melt-Based Processing Technologies: Advances and Perspectives
Previous Article in Journal
Synthesis and Comparative Study of Polyether-b-polybutadiene-b-polyether Triblock Copolymers for Use as Polyurethanes
Previous Article in Special Issue
Property Improvements of Silica-Filled Styrene Butadiene Rubber/Butadiene Rubber Blend Incorporated with Fatty-Acid-Containing Palm Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reactive Blending of Modified Thermoplastic Starch Chlorhexidine Gluconate and Poly(butylene succinate) Blending with Epoxy Compatibilizer

by
Nanthicha Thajai
1,
Pornchai Rachtanapun
2,3,4,
Sarinthip Thanakkasaranee
2,3,
Winita Punyodom
4,5,
Patnarin Worajittiphon
4,5,
Yuthana Phimolsiripol
2,3,
Noppol Leksawasdi
2,3,
Sukunya Ross
6,
Pensak Jantrawut
3,7 and
Kittisak Jantanasakulwong
2,3,4,*
1
Nanoscience and Nanotechnology (International Program/Interdisciplinary), Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
2
School of Agro-Industry, Faculty of Agro-Industry, Chiang Mai University, Mae-Hea, Mueang, Chiang Mai 50100, Thailand
3
Center of Excellent in Agro Bio-Circular-Green Industry (Agro BCG), Faculty of Agro-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
Department of Chemistry, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
6
Department of Chemistry, Faculty of Science, Naresuan University, Phitsanulok 65000, Thailand
7
Department of Pharmaceutical Sciences, Faculty of Pharmacy, Chiang Mai University, Muang, Chiang Mai 50200, Thailand
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(16), 3487; https://doi.org/10.3390/polym15163487
Submission received: 7 July 2023 / Revised: 10 August 2023 / Accepted: 16 August 2023 / Published: 21 August 2023
(This article belongs to the Special Issue Trendings in Biobased Polymers and Biocomposites)

Abstract

:
Biodegradable starch-based polymers were developed by melt-blending modified thermoplastic starch (MTPS) with poly(butylene succinate) (PBS) blended with epoxy resin (Er). A modified thermoplastic starch blend with chlorhexidine gluconate (MTPSCh) was prepared by melt-blending cassava starch with glycerol and chlorhexidine gluconate (CHG) 1.0% wt. The Er was melt-blended with PBS (PBSE) at concentrations of 0.50%, 1.0%, 2.5%, and 5.0% (wt%/wt%). The mechanical properties, water resistance, and morphology of the MTPSCh/PBSE blends were investigated. The MTPSCh/PBSE2.5% blend showed an improvement in tensile strength (8.1 MPa) and elongation at break (86%) compared to the TPSCh/PBS blend (2.6 MPa and 53%, respectively). In addition, water contact angle measurements indicated an increase in the hydrophobicity of the MTPSCh/PBSE blends. Thermogravimetric analysis showed an improvement in thermal stability when PBS was added to the MTPSCh blends. Fourier transform infrared spectroscopy data confirmed a new reaction between the amino groups of CHG in MTPSCh and the epoxy groups of Er in PBSE, which improved the interfacial adhesion of the MTPSCh/PBSE blends. This reaction improved the mechanical properties, water resistance, morphology, and thermal stability of the TPSCh/PBSE blends.

1. Introduction

Researchers have attempted to develop alternative materials for natural and biodegradable polymers from renewable resources to reduce the use of petroleum-derived plastics [1]. Biopolymers are major materials in the development of biodegradable film for packaging and agricultural and medical applications; they include polysaccharides, polypeptides, and polynucleotides. Polysaccharides are formed by the glycosidic linkage of monosaccharides. Starch, cellulose, chitosan, pectin, glycogen, and alginate are polysaccharide base materials. Biopolymers with an amide linkage are effective biomaterials for reactions with other reactive functional groups, such as keratin, sericin, fibroin, collagen, gelatin, and polypeptides [2]. Synthesis is a technique used to develop biopolymers via polymerization and the chemical modification of natural monomer materials that present inherent biodegradability, biocompatibility, and excellent mechanical properties. Among biopolymers, starch is one of the most promising candidates for the preparation of packaging applications. Starch has been used as a base film and filler due to its biodegradable ability, low cost, and renewability. Starch forms strong intermolecular hydrogen bonds and degrades before melting. It can be converted into thermoplastic starch (TPS) by gelatinization and plasticization. The preparation of TPS requires the incorporation of plasticizers such as sorbitol, glucose, and glycerol with high temperature and shear force. Glycerol (25–40%) is an effective plasticizer to develop TPS [3]. In particular, TPS has been extensively used for the development of bioplastic films and packaging because of its cost-effectiveness, complete biodegradability, and natural abundance [4]. TPS shows desirable properties, including the fact that it is colorless, odorless, and non-toxic [5]. However, TPS has several significant disadvantages, such as poor mechanical properties, low thermal processing ability, high water sensitivity, and plasticizer migration [6,7]. Several studies have investigated modified starch to improve the water sensitivity and reactivity of starch [8,9]. Modifications of starch are commonly conducted to alter its hydrophilicity by replacing the hydroxyl groups with chemically active groups. Cationic, anionic, and nonionic materials have been widely used for starch modification as they can be attached to the hydroxyl groups of starch. Modified cationic starches are widely used in the global paper industry, but few studies have investigated their application in preparing TPS [10,11,12,13]. It has been determined that the processability, water resistance, and mechanical properties of TPS can be improved by blending it with other completely biodegradable materials. To overcome the limitations of TPS, the blending of TPS with bioplastics, such as polylactic acid (PLA), poly (butylene succinate) (PBS), and poly(butylene adipate co-terephthalate) (PBAT), has been studied [14,15]. Reactive blending is an effective technique to improve biodegradable material properties [16]. TPS is usually blended with other natural materials, polymers, and nanoparticles, etc. [17]. Agricultural waste has been blended with TPS extensively due to its low cost, high added value, wide range of sources, and abundant materials. However, the incompatibility of the TPS blend provides poor morphology, interfacial tension, water resistance, and mechanical properties due to its structure and hydrophilic properties. It is important to develop TPS properties using starch modification, synthesis, composites, and reactive blending techniques. The development of new biopolymers should be targeted towards emphasizing degradation ability and environmental assessments.
In recent years, bioplastics research has been focused on property improvements, such as PLA, PBS, and PBAT. Increasing the amount of bioplastics development that utilizes renewable resources has become a top priority. However, high costs and the inadequacy of the mechanical properties prevent the use of bioplastics in a wide range of applications. Among these bioplastics, PLA is a biobased fully biodegradable polyester that is derived from agricultural materials. PLA shows high transparency, stiffness, tensile strength, biocompatibility, and processability. PLA has been limited because of its low degradation rate and high brittleness. PBAT is a ductile thermoplastic polyester with high mechanical properties and flexibility and is similar to polyethylene. PBAT has usually been used to toughen biopolymers with the addition of a reactive compatibilizer. Several studies have investigated a TPS blend with PBAT [18,19]. PBS is an aliphatic polyester that is polymerized from butanediol and succinic acid and can be derived from petroleum-based and bio-based renewable resources [20,21]. The mechanical properties and processability of PBS are similar to those of polyethylene. PBS is a commercial bioplastic that can be used to prepare plastic bags, films, and injection-molded packaging. The low temperature (130–160 °C) used to process PBS is lower than that for PLA. PBS has excellent properties, including good tensile strength (18–30 MPa), elongation (20–500%), ductility, biodegradability, and processing capabilities [22,23], compared with low-density polyethylene (tensile strength 10 MPa and elongation 500%) [24]. The use of bioplastics in the preparation of biopolymer blends is an effective method with which to develop the mechanical and water resistance properties of the blend. The addition of a compatibilizer is the key to improving the compatibility between the biopolymer and bioplastic blend.
Epoxy resin (Er) is a thermoset polymer. Its network structure is formed via a reaction between the epoxy groups of an epoxy monomer and the amino groups of a hardener. Er shows excellent mechanical properties, chemical resistance, low cost, and easy processing. Er is used as a matrix material for composites that are used in automotive and construction materials and in the electronics, coating, and aerospace fields [25]. The toughening of Er without reinforced materials is a major topic of research. Nanometal particle-reinforced Er has been extensively investigated [26,27]. Bisphenol A diglycidyl is an epoxy monomer part of Er. It is a reactive crosslinking agent that is usually used in polymers, such as polycarbonate [28], polycaprolactone [29], polyamine [30], and polylactic acid [31]. Er consists of reactive epoxide groups that crosslink with other functional groups, such as hydroxyl and carboxyl groups (–COOH), leading to improved morphology as well as the enhanced melting and crystallization properties of polymer blends [32,33]. However, the blending of Er and PBS has not been widely researched. In addition, the improvement of the mechanical, water resistance, and antimicrobial properties of starch blends modified with chlorhexidine gluconate (CHG), PBS, and Er has not been reported.
Therefore, the aim of this study was to develop MTPS from cassava starch blends with CHG, PBS, and Er to improve the mechanical properties, water resistance, morphology, thermal stability, and reactivity of the blends for medical, agricultural, and packaging applications. MTPS was used as the main matrix polymer because of its hydrophobicity and polarity, whereas CHG was mixed with MTPS to provide amino groups during the melt blending of the modified thermoplastic starch. Er was used to improve crosslinking and provide epoxy groups to PBSE. The reaction between the amino groups of MTPSCh and the epoxy groups of PBSE may induce interfacial crosslinking and improve the properties of the MTPSCh/PBSE blend.

2. Materials and Methods

2.1. Materials

Modified starch (modified tapioca starch, EXCELCAT®30) with a moisture content of 13 (wt%/wt%), DS of 0.028–0.032, and pH of 4.5–6.0 was purchased from Siam Modified Starch Co., Ltd., Pathum Thani, Thailand. Glycerol (99%) was purchased from Union Science Co. Ltd., Chiang Mai, Thailand. PBS pellets (BioPBS™ FD92PM, density 1.24 g/cm3, MFI 4 g/10 min at 190 °C) were purchased from PTT Global Chemical Pub Co., Ltd., Bangkok, Thailand. CHG (20% solution) was purchased from S. Tong Chemical Co., Ltd., Chiang Mai, Thailand. Er (diglycidyl ether of bisphenol, grade A 0302) was purchased from Easy Resin Co., Ltd., Nonthaburi, Thailand.

2.2. Sample Preparation

MTPS was prepared by mixing the modified starch with glycerol (70/30 wt%/wt%) and 500 mL of distilled water in a water bath (WNB 14 L0, Memmert GmbH + Co. KG, Bavaria, Germany) with an overhead stirrer at 70 °C for 30 min. CHG at 1.0 wt%/wt% was added to the MTPS as an additive to prepare MTPSCh. Distilled water in the premixed starch was evaporated in a hot air oven at 50 °C for 24 h. PBSE was prepared by melt-blending PBS with Er at 0.5%, 1.0%, 2.5%, and 5.0% (wt%/wt%) using a two-roll mill (PI-140, Pirom-Olarn Co. Ltd., Bangkok, Thailand) at 130 °C for 15 min. The melt blending of MTPSCh (90%) and PBSE (10%) were performed using a two-roll mill at 130 °C for 5 min. The samples were compressed into sheets using a hot compress at 130 °C for 5 min to evaluate their properties. The concentrations and codenames of the samples are listed in Table 1.

2.3. Tensile Properties

The tensile properties of the samples were measured following JISK-6251-7 using a tensile tester (Model H1KS, Hounsfield test equipment, Surrey, England) with a gap length of 10 mm and crosshead speed of 10 mm/min. The specimens were prepared as sheets with dimensions of a 5 mm width, 30 mm length, and 1 mm thickness. Five specimens from each of the samples were conditioned at 25 °C and 50 ± 2% RH for 24 h. The stress–strain curve, maximum force, and elongation at break were recorded.

2.4. Contact Angle

Water contact angles were measured using a drop shape analyzer (DSA30E, Krüss Co., Ltd., Hamburg, Germany). A droplet of water was dropped onto the surface of the sample, and images were recorded automatically every minute for 10 min. The samples were prepared as sheets (1 × 1 cm2) via hot-compression molding at 130 °C for 5 min. The surface of the hot-compressed sample was used to observe the water droplet contact angle. All the samples were conditioned at a relative humidity of 50 ± 2% RH for 24 h.

2.5. Scanning Electron Microscopy (SEM)

The morphology and fractured surfaces of the samples were analyzed using SEM (JSM-IT300LV model, JEOL Co., Ltd., Tokyo, Japan). The samples were broken in liquid nitrogen, evaporated in a hot air oven (60 °C) for 1 h, and then coated with a thin layer of gold. The fractured samples were observed at 15 kV under a vacuum.

2.6. Thermogravimetric Analysis (TGA)

TGA of all the samples was performed using a thermogravimetric analyzer (Mettler-Toledo STARe system TGA/DSC3+, Greifensee, Switzerland) at 10 °C/min under a nitrogen atmosphere from 25 to 550 °C. A sample weight of 10–15 mg was used to evaluate the degradation characteristics of the samples.

2.7. Reaction Mechanism

Fourier transform infrared (FTIR) spectra were measured to study the reaction mechanism of the samples using an FTIR spectrometer (FT/IR-4700, Jasco Corp., Tokyo, Japan) with a resolution of 4 cm−1 in the range of 600–4000 cm−1. The samples were prepared as films via hot-compression molding at 130 °C for 5 min.

2.8. Statistical Analysis

One-way ANOVA with SPSS software (SPSS version 17, Armonk, NY, USA) was used to analyze the results. The differences (p < 0.05) were estimated using the Duncan test.

3. Results and Discussion

3.1. Reaction Mechanism

FTIR is an important instrument in the study of the structures and functional group reactions in polymer blends. It was used to characterize the structures and reaction mechanisms of the MTPS and PBS blends with CHG and the Er compatibilizer addition. The FTIR spectra of the TPS, PBS, CHG, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends with 0.5–5.0% Er are shown in Figure 1. TPS and MTPS exhibited main peak bands at 3290, 1643, 1016, and 929 cm−1, which corresponded with the hydroxyl group (–OH), vibrations of the bending hydroxyl group (–OH) and the scissoring of two –CO stretching bands, respectively [34,35,36,37]. CHG presented characteristic peaks at 1580 (N–H), 1640 (C=N), 1095, and 1155 cm−1 (C–O–C of the chloride-bonded aromatic ring) [38,39,40]. In the FTIR spectrum of PBS, the characteristic peaks were observed at 2950, 1712, and 1162 cm−1, which correspond with the alkyl (–CH2), carbonyl (C=O), and (–CO) stretching vibrations of PBS, respectively [41,42]. The MTPSCh blend with PBS and PBSE exhibited the combination spectra of MTPS, CHG, PBS, and PBSE. The intensity of the –CO stretching main peak of MTPS at 1016 cm−1 was used to normalize the spectra of the blends. The MTPCh/PBS showed a shifting peak of C=O PBS at 1712 cm–1 to 1740 cm–1, which indicated the stretching of a new C=O vibration from the strong reaction between the carboxyl end groups (–COOH) of PBS and the amino groups (–NH) of CHG in MTPSCh. This reaction improved the elongation properties of MTPCh/PBS. MTPCh/PBSE showed a new peak at 1720 cm–1 and an overlap peak with increased intensity at 1720–1740 cm–1, which was attributed to the stretching of an ester carbonyl owing to the reactions of the epoxy groups in the PBSE and –COOH end groups of PBS with the amino groups of CHG in MTPSCh. This reaction improved the mechanical properties, water resistance, morphology, and thermal properties of the blends. The improvement in the properties of polymer blends owing to the reaction between the epoxy groups and amino groups was reported previously [36,43,44].

3.2. Mechanical Properties

The tensile properties of TPS, MTPS, MTPSCh, and MTPSCh melt-blended with PBS and PBSE at 0.5%, 1.0%, 2.5%, and 5.0% of Er are presented in Figure 2. The tensile modulus was calculated from the slope of the stress–strain curve. The tensile moduli the of TPS, MTPS, and MTPSCh samples were 12, 105, and 337 MPa, while the MTPSCh/PBSE blends with the Er 0, 0.5, 1, 2.5, and 5% samples were 26, 118, 85, 109, and 27 MPa, respectively. MTPSCh showed the highest tensile modulus due to the high crosslinking between MTPS and CHG. The tensile strength of MTPS (10.2 MPa) was higher than that of TPS (3.3 MPa). However, the elongation at break of MTPS (1.5%) was lower than that of TPS (54.8%). Upon adding CHG 1.0% to MTPS, the tensile strength increased to 15.7 MPa, and the elongation at break decreased to 6.7% owing to the formation of a crosslinking network of starch with CHG. The reaction-induced crosslinking of TPS after the CHG addition was reported previously [38]. The MTPSCh/PBS blend had a reduced tensile strength of 2.6 MPa and an increased elongation at break of 53.1%, owing to the flexibility of the PBS ductile polymer [22]. The tensile strength and the elongation at break of the MTPSCh/PBSE blends with 0.5–2.5% Er were higher compared to those of the MTPSCh/PBS blends. The increase in the tensile modulus, tensile strength, and elongation at break of the MTPSCh/PBSE samples was because of the interfacial crosslinking network between the CHG in MTPSCh and the epoxy groups in PBSE, the plasticization effect of Er, and the increase in interfacial compatibility between PBSE and MTPSCh. The formation of a crosslinking network because of the reaction between the epoxy group of Er with –NH of CHG and the plasticization effect of Er was reported previously [31,39]. However, the tensile properties of the TPSCh/PBSE5 blend decreased because of the excessive amount of Er which acted as a high plasticizer content.

3.3. Contact Angle

The water contact angle is an indicator of surface tension, the hydrophilic and hydrophobic properties, and the water resistance of a material’s surface. The water contact angles of the TPS, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends with 0.5%, 1.0%, 2.5%, and 5.0% Er are shown in Figure 3. A water droplet was absorbed onto the surface, and the contact angle was automatically recorded at 0, 2, 4, 6, 8, and 10 min after its absorption. TPS had the lowest contact angle after 10 min (8.5°), whereas MTPS showed at 23.5°. TPS exhibited higher hydrophilicity than MTPS. The addition of CHG to MTPS increased the contact angle at 10 min to 38.1° because of a reaction occurring in the crosslinking network of the MTPSCh blend. The crosslinking with TPS improved the water contact angle [34,38,45,46]. In the MTPSCh blend with PBS and PBSE at 0.5–5.0% Er, the water contact angle at 10 min increased to 43.5° (MTPSCh/PBS), 44.1° (MTPSCh/PBSE0.5), 45.8° (MTPSCh/PBSE1), 45.4° (MTPSCh/PBSE2.5), and 45.3° (MTPSCh/PBSE5), as shown in Figure 3b. The water contact angle was improved by the addition of PBS and PBSE to the MTPSCh blends. The addition of Er to PBS rarely improved the water resistance of the TPSCh/PBSE blends. The contact angles of polyethylene (67°), polypropylene (105°), polyethylene terepthalate (91°), and nylon (73°) have been reported [47,48]. The results indicated that the addition of PBS to TPSCh increased the interfacial tension of the MTPSCh/PBS blend, owing to the hydrophobicity of the PBS polymer. Er addition to PBSE slightly increases the contact angle and interfacial tension because of the occurrence of the crosslinking reaction via Er [43,49,50].

3.4. Scanning Electron Microscopy (SEM)

The microstructure and morphology of the polymer blend are closely related to the mechanical properties, processing ability, rheology, thermal properties, and water resistance. SEM was used to observe the microstructure of the MTPS and PBS blend. Figure 4 shows the SEM images of the fracture surfaces of MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE with 0.5–5.0% Er blends. MTPS showed a smooth fracture surface image, while MTPSCh exhibited an increase in the rough fracture surface because of the agglomeration of the crosslink phase between CHG and starch or glycerol [38]. The MTPSCh/PBS blend presented large PBS particles (10–15 μm) scattered in the MTPSCh, but the MTPSCh/PBSE with 0.5–2.5% Er had smaller white particle sizes (~1 μm). The small PBS particles indicated high interfacial adhesion between MTPSCh and PBSE owing to the addition of Er, the occurrence of the interfacial reaction, and a high shear rate of mixing [31,41]. The MTPSCh/PBSE2.5 exhibited small particles (~1 μm) with a smooth fracture surface, owing to the high interfacial adhesion between TPSCh and PBSE. However, excessive amounts of Er caused the particle size of MTPCh/PBSE5 to slightly increase. The high interfacial adhesion between MTPSCh and PBS via the reaction with Er resulted in a small particle size and smooth fracture surface morphology, which improved the tensile and water resistance properties of the blend.

3.5. Thermal Stability

The thermal gravimetric analysis was used to study phase decomposition, including various hydrates during heating. The thermal stabilities of PBS, TPS, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE with 0.5–5.0% Er were studied using TGA, as shown in Figure 5. The thermal decomposition of PBS presented one stage in the thermogram at approximately 400 °C. The thermal decomposition of the TPS, MTPS, and MTPSCh blends showed two stages in the TGA thermograms. MTPS was a hydrophilic material which presented an –OH peak (3290 cm−1) in the FTIR spectra (Figure 1). The –OH groups of MTPS formed hydrogen bonding with moisture, holding water inside its structure. The first stage occurred during the evaporation of free water at approximately 100 °C by the water absorption of MTPS. The second stage was at 260–300 °C, which indicated the decomposition stage of starch [23]. The first and second decomposition stages of TPS showed the lowest weight loss owing to the low thermal stability of TPS. The MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends showed less weight loss in the first stage compared to MTPS because of the decomposition of CHG of the blends. The MTPSCh samples with PBS and PBSE at 0.50%, 1.0%, 2.5%, and 5.0% showed thermal decomposition in three steps at 100–160 °C, 260–300 °C, and 320–380 °C owing to the evaporation of free water and CHG, the decomposition of starch, and the decomposition of Er and PBS in the samples, respectively [51]. The high weight loss% of the MTPSCh/PBS and MTPS/PBSE blends at 330 °C indicated a high thermal stability of PBS and the occurrence of crosslinking via the Er reaction. The MTPSCh/PBSE2.5 blend had the highest weight loss at approximately 330 °C because of the suitable crosslinking via the reaction with Er. TPS presented a low ash content at 330–530 °C due to the high degradation rate of the starch material at a high temperature, while PBS showed the lowest ash content from the degradation of ester groups in the PBS structure. MTPSCh showed more ash value than MTPS owing to the crosslinking of MTPS via the CHG reaction. The reaction between TPS and CHG was reported previously [35]. The MTPSCh/PBS and MTPSCh/PBSE samples exhibited almost the same ash content, which indicated the combination of the high degradation rate of PBS and the remaining crosslinking part inside MTPSCh.

4. Conclusions

Biodegradable starch-based blends were successfully developed through the melt blending of MTPSCh and PBS with Er. FTIR spectroscopy confirmed the reaction between the amino groups of CHG in MTPSCh and the epoxy groups of Er in PBSE. The tensile strength, elongation at break, morphology, water resistance, and thermal properties of the MTPSCh/PBSE blends were improved compared to TPS, especially at 2.5 wt% Er. Because of the high interfacial adhesion owing to the reaction and interaction of the blends, the MTPSCh/PBSE blends exhibited a morphology with small PBS particles distributed in an MTPS matrix. The addition of PBS and PBSE to the MTPSCh blends improved their thermal decomposition behavior. The reaction that occurred indicated an improvement of the compatibility of the MTPSCh/PBSE blend compared to MTPSCh/PBS, which improved the mechanical, water resistance, and thermal properties of the blends. Biodegradable polymers of MTPSCh/PBSE blends with good mechanical properties, thermal stability, and water resistance can be used in medical, agricultural, and packaging applications.

Author Contributions

Methodology, N.T., W.P. and K.J.; Validation, K.J.; Formal analysis, N.T.; Investigation, N.T.; Resources, W.P. and K.J.; Data curation, N.T.; Writing—original draft, N.T.; Writing—review & editing, P.R., S.T., W.P., P.W., Y.P., N.L., S.R., P.J. and K.J.; Visualization, P.R., S.T., W.P., P.W., Y.P., N.L., S.R., P.J. and K.J.; Supervision, P.R., S.T., W.P., P.W., S.R. and K.J.; Project administration, K.J.; Funding acquisition, W.P. and K.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research project was supported by Fundamental Fund 2023, Chiang Mai University. This research work was partially supported by 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 upon request from the corresponding author.

Acknowledgments

The authors gratefully acknowledge the Faculty of Agro-Industry, Chiang Mai University, for their support. This research project was supported by Fundamental Fund 2023, Chiang Mai University. This research work was partially supported by Chiang Mai University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CHG, chlorhexidine gluconate; Er, epoxy resin; FTIR, Fourier transform infrared spectroscopy; MTPS, modified thermoplastic starch; MTPSCh, modified thermoplastic starch blend with chlorhexidine gluconate; PBS, poly(butylene succinate); PBSE, the Er was melt-blended with PBS; SEM, scanning electron microscopy; TGA, thermogravimetric analysis.

References

  1. Gross, R.A.; Kalra, B. Biodegradable polymers for the environment. Science 2002, 297, 803–807. [Google Scholar] [CrossRef] [PubMed]
  2. Yakul, K.; Kaewsalud, T.; Techapun, C.; Seesuriyachan, P.; Jantanasakulwong, K.; Watanabe, M.; Takenaka, S.; Chaiyaso, T. Enzymatic valorization process of yellow cocoon waste for production of antioxidative sericin and fibroin film. J. Chem. Technol. Biotechnol. 2021, 96, 953–962. [Google Scholar] [CrossRef]
  3. Dang, K.M.; Yoksan, R.; Pollet, E.; Averous, L. Morphology and properties of thermoplastic starch blended with biodegradable polyester and filled with halloysite nanoclay. Carbohydr. Polym. 2020, 242, 116392. [Google Scholar] [CrossRef] [PubMed]
  4. Khan, B.; Niazi, M.B.K.; Samin, G.; Jahan, Z. Thermoplastic starch: A possible biodegradable food packaging material—A review. J. Food Process Eng. 2017, 40, e12447. [Google Scholar] [CrossRef]
  5. Qin, Y.; Liu, Y.; Zhang, X.; Liu, J. Development of active and intelligent packaging by incorporating betalains from red pitaya (Hylocereus polyrhizus) peel into starch/polyvinyl alcohol films. Food Hydrocoll. 2020, 100, 105410. [Google Scholar] [CrossRef]
  6. Roz, A.; Carvalho, A.; Gandini, A.; Curvelo, A. The effect of plasticizers on thermoplastic starch compositions obtained by melt processing. Carbohydr. Polym. 2006, 63, 417–424. [Google Scholar] [CrossRef]
  7. Ma, X.; Yu, J. The plastcizers containing amide groups for thermoplastic starch. Carbohydr. Polym. 2004, 57, 197–203. [Google Scholar] [CrossRef]
  8. Namazi, H.; Fathi, F.; Dadkhah, A. Hydrophobically modified starch using long-chain fatty acids for preparation of nanosized starch particles. Sci. Iran. 2011, 18, 439–445. [Google Scholar] [CrossRef]
  9. Samu, R.; Moulee, A.; Kumar, V.G. Effect of charge and hydrophobicity on adsorption of modified starches on polyester. J. Colloid Interface Sci. 1999, 220, 260–268. [Google Scholar] [CrossRef]
  10. Chaisawang, M.; Suphantharika, M. Effects of guar gum and xanthan gum additions on physical and rheological properties of cationic tapioca starch. Carbohydr. Polym. 2005, 61, 288–295. [Google Scholar] [CrossRef]
  11. Guan, Y.; Qian, L.; Xiao, H.; Zheng, A.; He, B. Synthesis of a novel antimicrobial-modified starch and its adsorption on cellulose fibers: Part II—Adsorption behaviors of cationic starch on cellulose fibers. Cellulose 2008, 15, 619–629. [Google Scholar] [CrossRef]
  12. Putro, J.N.; Ismadji, S.; Gunarto, C.; Soetaredjo, F.E.; Ju, Y.H. A study of anionic, cationic, and nonionic surfactants modified starch nanoparticles for hydrophobic drug loading and release. J. Mol. Liq. 2020, 298, 112034. [Google Scholar] [CrossRef]
  13. Yang, H.; Qiu, L.; Qian, X.; Shen, J. Filler modification for papermaking with cationic starch and carboxymethyl cellulose: A comparative study. BioResources 2013, 8, 5449–5460. [Google Scholar] [CrossRef]
  14. Lorena, H.; Pulgarin, C.; Caicedo, C.; Lopez, E.F. Effect of surfactant content on rheological, thermal, morphological and surface properties of thermoplastic starch (TPS) and polylactic acid (PLA) blends. Heliyon 2022, 8, e10833. [Google Scholar]
  15. Beluci, N.C.L.; Santos, J.D.; Carvalho, F.A.; Yamashita, F. Reactive biodegradable extruded blends of thermoplastic starch and polyesters. Carbohydr. Polym. Technol. Appl. 2023, 5, 100274. [Google Scholar] [CrossRef]
  16. Jantanasakulwong, K.; Kobayashi, Y.; Keiichi, K.; Toshiaki, O. Thermoplastic vulcanizate based on poly(lactic acid) and acrylic rubber blended with ethylene ionomer. J. Macromol. Sci. Phys. 2016, 55, 1068–1085. [Google Scholar] [CrossRef]
  17. Cheng, H.; Chen, L.; McClements, D.J.; Yang, T.; Zhang, Z.; Ren, F.; Miao, M.; Tian, Y.; Jin, Z. Starch-based biodegradable packaging materials: A review of their preparation, characterization and diverse applications in the food industry. Trends Food Sci. Technol. 2021, 114, 70–82. [Google Scholar] [CrossRef]
  18. Lendvai, L.; Apostolov, A.; Kocsis, J.K. Characterization of layered silicate reinforced blended of thermoplastic starch (TPS) and poly(butylene adipate-co-terephthalate). Carbohydr. Polym. 2017, 173, 566–572. [Google Scholar] [CrossRef]
  19. Wei, D.; Wang, H.; Xiao, H.; Zheng, A.; Yang, Y. Morphology and mechanical properties of poly(butylene adipate-co-terephthalate)/potato starch blends in the presence of synthesized reactive compatibilizer or modified poly(butylene adipate-co-terephthalate). Carbohydr. Polym. 2015, 123, 275–282. [Google Scholar] [CrossRef]
  20. Liu, L.; Yu, J.; Cheng, L.; Yang, X. Biodegradability of poly (butylene succinate) (PBS) composite reinforced with jute fibre. Polym. Degrad. Stab. 2009, 94, 90–94. [Google Scholar] [CrossRef]
  21. Yun, I.S.; Hwang, S.W.; Shim, J.K.; Seo, K.H. A study on the thermal and mechanical properties of poly (butylene succinate)/thermoplastic starch binary blends. Int. J. Precis. Eng. Manuf. Green Technol. 2016, 3, 289–296. [Google Scholar] [CrossRef]
  22. Ayu, R.S.; Khalina, A.; Harmaen, A.S.; Zaman, K.; Jawaid, M.; Lee, C.H. Effect of modified tapioca starch on mechanical, thermal, and morphological properties of PBS blends for food packaging. Polymers 2018, 10, 1187. [Google Scholar] [CrossRef] [PubMed]
  23. Li, J.; Luo, X.; Lin, X.; Zhou, Y. Comparative study on the blends of PBS/thermoplastic starch prepared from waxy and normal corn starches. Starch-Stärke 2013, 65, 831–839. [Google Scholar] [CrossRef]
  24. Szlachetka, O.; Witkowska-Dobrev, J.; Baryla, A.; Dohojda, M. Low-density polyethylene (LDPE) building films—Tensile properties and surface morphology. J. Build. Eng. 2021, 44, 103386. [Google Scholar] [CrossRef]
  25. Njuguna, J.; Pielichowski, K.; Alcock, J.R. Epoxy-based fibre reinforced nanocomposites. Adv. Eng. Mater. 2007, 9, 835–847. [Google Scholar] [CrossRef]
  26. Komarneni, S. Nanocomposites. J. Mater. Chem. 1992, 2, 1219–1230. [Google Scholar] [CrossRef]
  27. Rajsekhar, V.; Gattu, M. Size-effect testing: Nano-alumina enhances fracture toughness of epoxy resins. Theor. Appl. Fract. Mech. 2023, 125, 103859. [Google Scholar] [CrossRef]
  28. Wu, D.; Wu, L.; Zhang, M.; Wu, L. Effect of epoxy resin on rheology of polycarbonate/clay nanocomposites. Eur. Polym. J. 2007, 43, 1635–1644. [Google Scholar] [CrossRef]
  29. Könczöl, L.; Döll, W.; Buchholz, U.; Mülhaupt, R. Ultimate properties of epoxy resins modified with a polysiloxane–polycaprolactone block copolymer. J. Appl. Polym. Sci. 1994, 54, 815–826. [Google Scholar] [CrossRef]
  30. 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. B Polym. Phys. 2000, 38, 23–33. [Google Scholar] [CrossRef]
  31. Kiatiporntipthak, 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. [Google Scholar] [CrossRef] [PubMed]
  32. Verma, C.; Olasunkanmi, L.O.; Akpan, E.D.; Quraishi, M.A.; Dagdag, O.; El Gouri, M.; Sherif, E.S.M.; Ebenso, E.E. Epoxy resins as anticorrosive polymeric materials: A review. React. Funct. Polym. 2020, 156, 104741. [Google Scholar] [CrossRef]
  33. 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]
  34. Jantanasakulwong, K.; Leksawasdi, N.; Seesuriyachan, P.; Wongsuriyasak, S.; Techapun, C.; Ougizawa, T. Reactive blending of thermoplastic starch and polyethylene-graft-maleic anhydride with chitosan as compatibilizer. Carbohydr. Polym. 2016, 153, 89–95. [Google Scholar] [CrossRef]
  35. Deeyai, P.; Suphantharika, M.; Wongsagonsup, R.; Dangtip, S. Characterization of modified tapioca starch in atmospheric argon plasma under diverse humidity by FTIR spectroscopy. Chin. Phys. Lett. 2013, 30, 018103. [Google Scholar] [CrossRef]
  36. Mendes, J.F.; Paschoalin, R.T.; Carmona, V.B.; Sena Neto, A.R.S.; Marques, A.C.P.; Marconcini, J.M.; Mattoso, L.H.C.; Medeiros, E.S.; Oliveira, J.E. Biodegradable polymer blends based on corn starch and thermoplastic chitosan processed by extrusion. Carbohydr. Polym. 2016, 137, 452–458. [Google Scholar] [CrossRef]
  37. Jantanasakulwong, K.; Wongsuriyasak, S.; Rachtanapun, P.; Seesuriyachan, P.; Chaiyaso, T.; Leksawasdi, N.; Techapun, C. Mechanical properties improvement of thermoplastic corn starch and polyethylene of carboxymethyl cellulose. Int. J. Biol. Macromol. 2018, 120, 297–301. [Google Scholar] [CrossRef]
  38. Thajai, N.; Jantanasakulwong, K.; Rachtanapun, P.; Jantrawut, P.; Kiattipornpithak, K.; Kanthiya, T.; Punyodom, W. Effect of chlorhexidine gluconate on mechanical and anti-microbial properties of thermoplastic cassava starch. Carbohydr. Polym. 2022, 275, 118690. [Google Scholar] [CrossRef]
  39. 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 bio-thermoplastic elastomer made from thermoplastic starch-chitosan blended with epoxidized natural rubber. Carbohydr. Polym. 2020, 242, 116421. [Google Scholar] [CrossRef]
  40. Raso, E.M.G.; Cortes, M.E.; Teixeira, K.I.; Franco, M.B.; Mohallem, N.D.S.; Sinisterra, R.D. A new controlled release system of chlorhexidine and chlorhexidine: Βcd inclusion compounds based on porous silica. J. Incl. Phenom. Macrocycl. Chem. 2010, 67, 159–168. [Google Scholar] [CrossRef]
  41. Yin, Q.; Chen, F.; Zhang, H.; Liu, C. Fabrication and characterisation of thermoplastic starch/poly (butylene succinate) blends with maleated poly (butylene succinate) as compatibilizer. Plast. Rubber Compos. 2015, 44, 362–367. [Google Scholar] [CrossRef]
  42. Han, X.; Jin, Y.; Wang, B.; Tian, H.; Weng, Y. Reinforcing and toughening modification of PPC/PBS blends Compatibilized with epoxy terminated hyperbranched polymers. J. Polym. Environ. 2022, 30, 461–471. [Google Scholar] [CrossRef]
  43. Yan, Y.; Dou, Q. Effect of peroxide on compatibility, microstructure, rheology, crystallization, and mechanical properties of PBS/Waxy starch composites. Starch-Stärke 2021, 73, 2000184. [Google Scholar] [CrossRef]
  44. Hong, L.F.; Cheng, L.H.; Lee, C.Y.; Peh, K.K. Characterisation of physicochemical properties of propionylated corn starch and its application as stabilizer. Food Technol. Biotechnol. 2015, 53, 278–285. [Google Scholar] [CrossRef] [PubMed]
  45. Jantanasakulwong, K.; Leksawasdi, N.; Seesuriyachan, P.; Wongsuriyasak, S.; Techapun, C.; Ougizawa, T. Reactive blending of thermoplastic starch, epoxidized natural rubber and chitosan. Eur. Polym. J. 2016, 84, 292–299. [Google Scholar] [CrossRef]
  46. Zhou, J.; Zhang, J.; Ma, Y.; Tong, J. Surface photo-crosslinking of corn starch sheets. Carbohydr. Polym. 2008, 74, 405–410. [Google Scholar] [CrossRef]
  47. Lazrak, C.; Kabouchi, B.; Hammi, M.; Famiri, A.; Ziani, M. Structural study of maritime pine wood and recycled high-density poltethylene (HDPEr) plastic composite using Infrared-ATR spectroscopy, X-ray diffraction, SEM and contact angle measurements. Case Stud. Constr. Mater. 2019, 10, e00227. [Google Scholar]
  48. Choi, H.-J.; Kim, M.S.; Ahn, D.; Yeo, S.Y.; Lee, S. Electrical percolation threshold of carbon black in a polymer matrix and its application to antistatic fibre. Sci. Rep. 2019, 9, 6338. [Google Scholar] [CrossRef]
  49. Soccio, M.; Dominici, F.; Quattrosoldi, S.; Luzi, F.; Munari, A.; Torre, L.; Lotti, N.; Puglia, D. PBS-based green copolymer as an efficient compatibilizer in thermoplastic inedible wheat flour/poly (butylene succinate) blends. Biomacromolecules 2020, 21, 3254–3269. [Google Scholar] [CrossRef]
  50. Zeng, J.B.; Jiao, L.; Li, Y.D.; Srinivasan, M.; Li, T.; Wang, Y.Z. Bio-based blends of starch and poly (butylene succinate) with improved miscibility, mechanical properties, and reduced water absorption. Carbohydr. Polym. 2011, 83, 762–768. [Google Scholar] [CrossRef]
  51. Xu, J.; Chen, Y.; Tian, Y.; Yang, Z.; Zhao, Z.; Du, W.; Zhang, X. Effect of ionic liquid 1-buyl-3-methylimidazolium halide on the structure and tensile property of PBS/corn starch blends. Int. J. Biol. Macromol. 2021, 172, 170–177. [Google Scholar] [CrossRef] [PubMed]
Figure 1. FTIR spectra of TPS, PBS, CHG, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends with Er at 0.5–5.0%.
Figure 1. FTIR spectra of TPS, PBS, CHG, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends with Er at 0.5–5.0%.
Polymers 15 03487 g001
Figure 2. Tensile properties of TPS, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends with 0.5–5.0% Er. Mean values of the elongation at break (lowercase letters) and maximum tensile strength (uppercase letters) differ significantly (p < 0.05).
Figure 2. Tensile properties of TPS, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBSE blends with 0.5–5.0% Er. Mean values of the elongation at break (lowercase letters) and maximum tensile strength (uppercase letters) differ significantly (p < 0.05).
Polymers 15 03487 g002
Figure 3. Water contact angle of TPS, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBS with 0.5%, 1.0%, 2.5%, and 5.0% of Er; (a) water contact angle of samples and (b) image of water contact angle at 10 min.
Figure 3. Water contact angle of TPS, MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBS with 0.5%, 1.0%, 2.5%, and 5.0% of Er; (a) water contact angle of samples and (b) image of water contact angle at 10 min.
Polymers 15 03487 g003
Figure 4. SEM images of MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBS with 0.5%, 1.0%, 2.5%, and 5.0% Er.
Figure 4. SEM images of MTPS, MTPSCh, MTPSCh/PBS, and MTPSCh/PBS with 0.5%, 1.0%, 2.5%, and 5.0% Er.
Polymers 15 03487 g004
Figure 5. Thermograms of PBS, TPS, MTPS, MTPSCH1, MTPSCH1/PBS, and MTPSCH1/PBS with Er at 0.5%, 1.0%, 2.5%, and 5.0%.
Figure 5. Thermograms of PBS, TPS, MTPS, MTPSCH1, MTPSCH1/PBS, and MTPSCH1/PBS with Er at 0.5%, 1.0%, 2.5%, and 5.0%.
Polymers 15 03487 g005
Table 1. Code names and compositions of modified thermoplastic starch blend with chlorhexidine gluconate (MTPSCh), poly (butylene succinate) (PBS), and epoxy resin (Er) blends.
Table 1. Code names and compositions of modified thermoplastic starch blend with chlorhexidine gluconate (MTPSCh), poly (butylene succinate) (PBS), and epoxy resin (Er) blends.
Samples Composition (wt%/wt%)
MTPSChPBSEr
PBS-100-
MTPSCh100--
MTPSCh/PBS9010-
MTPSCh/PBSE0.5909.950.05
MTPSCh/PBSE1909.900.10
MTPSCh/PBSE2.5909.750.25
MTPSCh/PBSE5909.500.50
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Thajai, N.; Rachtanapun, P.; Thanakkasaranee, S.; Punyodom, W.; Worajittiphon, P.; Phimolsiripol, Y.; Leksawasdi, N.; Ross, S.; Jantrawut, P.; Jantanasakulwong, K. Reactive Blending of Modified Thermoplastic Starch Chlorhexidine Gluconate and Poly(butylene succinate) Blending with Epoxy Compatibilizer. Polymers 2023, 15, 3487. https://doi.org/10.3390/polym15163487

AMA Style

Thajai N, Rachtanapun P, Thanakkasaranee S, Punyodom W, Worajittiphon P, Phimolsiripol Y, Leksawasdi N, Ross S, Jantrawut P, Jantanasakulwong K. Reactive Blending of Modified Thermoplastic Starch Chlorhexidine Gluconate and Poly(butylene succinate) Blending with Epoxy Compatibilizer. Polymers. 2023; 15(16):3487. https://doi.org/10.3390/polym15163487

Chicago/Turabian Style

Thajai, Nanthicha, Pornchai Rachtanapun, Sarinthip Thanakkasaranee, Winita Punyodom, Patnarin Worajittiphon, Yuthana Phimolsiripol, Noppol Leksawasdi, Sukunya Ross, Pensak Jantrawut, and Kittisak Jantanasakulwong. 2023. "Reactive Blending of Modified Thermoplastic Starch Chlorhexidine Gluconate and Poly(butylene succinate) Blending with Epoxy Compatibilizer" Polymers 15, no. 16: 3487. https://doi.org/10.3390/polym15163487

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