DOI:
10.1039/D5GC01294E
(Critical Review)
Green Chem., 2025, Advance Article
Recent advances in biodegradable polymer blends and their biocomposites: a comprehensive review
Received
13th March 2025
, Accepted 14th July 2025
First published on 10th September 2025
Abstract
The growing environmental concerns over plastic pollution and sustainability have led to increased interest in biodegradable polymers as alternatives to conventional plastics. This concern has led to the United Nations resolution of March 2022 calling for urgent action to eradicate plastic pollution globally by 2040 as more than 90% of the global plastic production from 2018 to 2022 was fossil-based, significantly contributing to plastic pollution. In response, there has been a growing shift towards sustainable materials, with biodegradable polymers emerging as a critical solution to mitigate the environmental impacts. However, the properties of biodegradable polymers are at variance with conventional fossil-based plastics in many applications. One way to solve this problem is to re-engineer their properties through polymer blending, a strategy that combines the properties of two or more polymers, aided by compatibilization to improve polymer miscibility and properties. While numerous reviews have focused on biodegradable polymer blends, this article offers a unique contribution by comprehensively examining both biodegradable polymer blends and their reinforced biocomposites within a single review, an area that has seen limited coverage in recent years. This review discusses recent advancements in biodegradable polymer blends and reinforced biocomposites, focusing on material properties, compatibilization techniques, and environmental impact. Key biodegradable polymer blends and reinforced biocomposites based on polylactic acid (PLA), polyhydroxyalkanoates (PHAs), polybutylene succinate (PBS), Polybutylene adipate terephthalate (PBAT), and thermoplastic starch (TPS) are discussed, with a focus on miscibility, compatibilization and the effects on properties. It was found that compatibilizers such as maleic anhydride, dicumyl peroxide, and Joncryl play significant roles in polymer blend miscibility kinetics and compatibility while fillers such as turmeric, cinnamon, coffee ground powder, and rice straw have contributed to improving the mechanical properties and biodegradability of composites. This combined approach of blending and filler reinforcement represents a critical innovation for producing high-performance biodegradable materials. The review examines applications in packaging, agriculture, and biomedical fields, along with the environmental impacts of these materials, such as their biodegradation pathways and ecotoxicity. Lastly, the review discusses future outlooks, including potential breakthroughs and integrating biodegradable polymers into the circular economy.
Green foundation
1. The manuscript discusses recent developments in biodegradable polymer blends and their biocomposites, focusing on improving mechanical, thermal, and biodegradation properties. Innovations such as nanocomposite reinforcement, compatibilization techniques, and blending biodegradable polymers with bio-based fillers enhance material performance, enhancing environmental sustainability.
2. With increasing global plastic waste concerns, biodegradable polymer blends offer sustainable alternatives to fossil-based plastics. This review is significant for industries like packaging, agriculture, and biomedicine, addressing limitations of biodegradable polymers while enhancing their processability and eco-friendliness. It aligns with circular economy principles and global policies on pollution reduction.
3. Future research will explore advanced compatibilizers, reinforcements, and process optimization for biodegradable blends. This study provides critical insights into designing sustainable materials, helping shape green chemistry by fostering innovation in eco-friendly polymer development.
|
1. Introduction
The increasing urgency to address global plastic pollution and environmental sustainability has brought biodegradable polymers to the forefront of materials research and innovation. These polymers, derived from renewable sources, offer the potential to reduce dependence on fossil-based plastics and their associated ecological impact. However, their performance limitations in terms of mechanical strength, thermal stability, and processability restrict their broader application. Addressing these limitations through polymer blending and reinforcement with biocompatible fillers has opened new avenues in biodegradable material development.
The aim of this review article is to discuss the recent advancements in biodegradable polymer blends and reinforced biocomposites, focusing on miscibility, compatibilization, and their effects on properties. The review also examines biodegradation and ecotoxicity of blends and their biocomposites, along with their applications in food packaging, agriculture, and biomedical fields. Lastly, the review discusses the environmental impact assessment and integration of biodegradable polymers into the circular economy. While numerous reviews have focused on biodegradable polymer blends, this article offers a unique contribution by comprehensively examining both biodegradable polymer blends and their reinforced composites within a single review, an area that has seen limited coverage in the last five years (2019–2024).
The review article is structured as follows:
Section 2 gives the general background, section 3 introduces the fundamentals of biodegradable polymer blending, including key design considerations, section 4 discusses recent developments in blends based on PLA, PBAT, PHAs, TPS, PBS, and related systems, emphasizing miscibility and compatibility. This section also discusses the role played by natural and inorganic fillers in enhancing blend performance for the above blend systems. Section 5 examines the biodegradation behaviour of blends and composites, and section 6 covers application areas such as biomedical, packaging, and agriculture. The environmental impact assessment is discussed in section 7, section 8 presents the summary of key findings, while the last section presents the outlined future directions, including integration into circular economy frameworks.
2. General background
Biodegradable polymers are continuously in the research spotlight owing to their perceived benefits compared with non-biodegradable polymers. Growing environmental awareness has driven a shift towards developing biodegradable materials from renewable sources to replace conventional, non-biodegradable plastics, especially in packaging.1–4 The production of bio-degradable polymers is still low compared with fossil-based polymers; however, increasing awareness is driving global production, with significant increase in the last 5 years (Fig. 1). According to the 2023 Plastics Europe report,5 global plastic production grew from about 370 million tons (Mt) in 2018 to 400 Mt in 2022. Of these, fossil-based plastic grew from 339.4 Mt to 382, accounting for more than 90% of the global production and bio-based plastics grew from 1.1 Mt to 2.3 Mt, accounting for about 0.5% by 20226 (Fig. 1a and b). The plastics industry is increasingly interested in biopolymers, which can reduce carbon emissions by substituting fossil-based polymers.7,8 Hence there is a need to increase the production of bio-based polymers to replace fossil-based plastic in different applications. While bio-based polymers are environmentally friendly, their higher cost and performance limitations, such as brittleness and low mechanical performance, have hindered their widespread adoption.9 Consequently, the scientific community is focusing on tailoring their properties to performance and market demands.10,11 One key approach is to re-engineer their properties through a blend of biodegradable polymers and sometimes other materials to enhance their properties which have been demonstrated in many polymer blends.12–14 Blends of biodegradable polymers are gradually gaining attention as a critical approach for addressing the drawbacks of single biodegradable polymers, especially in areas such as mechanical properties, processability, and degradation rates. For instance, the blend of a slow-degrading polymer with a fast-degrading one can led to a material that degrades at a rate that benefits the target application.15,16 Blending offers a cost-effective and efficient way to tailor characteristics while adhering to low carbon footprint regulations.17,18 By combining two or more biodegradable polymers, scientists can create materials that possess a balance of desired characteristics that single polymers may lack.19 For example, polylactic acid (PLA), which is known for its mechanical strength and transparency, is often brittle and lacks flexibility. The incorporation of flexible polymers like polyhydroxyalkanoates (PHAs) and polybutylene adipate terephthalate (PBAT) can enhance the flexibility of the material without compromising its strength, making it suitable for a wider range of applications. Blending biodegradable polymers with non-degradable polymers enhances mechanical and functional properties while introducing partial biodegradability, reducing environmental impact compared with entirely non-degradable materials. Additionally, in biodegradable polymer blends, the degradation rates can be controlled especially in applications such as medical implants or agricultural films where the material's lifetime is a factor to consider. This ability to control properties through blending makes biodegradable polymer blends a critical component in the development of sustainable materials.
 |
| Fig. 1 (a) Global plastic production from 2018 to 2022, (b) global plastic production by polymer type from 2018 to 2022. Data adapted from Plastics Europe 2023 report https://plasticseurope.org/knowledge-hub/plastics-the-fast-facts-2023/ accessed January 4, 2025. | |
Also, the need to blend biodegradable polymers arises from the necessity to enhance processability and cost reduction. However, certain biodegradable polymers, although they have good properties, have some drawbacks in terms of processing. For example, they may need low temperatures, have short processing windows, or are difficult to form. When combined with polymers that have improved processing properties, the overall processing of these materials can be made easier and cheaper. Blending biodegradable polymers to make binary, ternary, quaternary blends can balance cost, processability, and performance resulting from the complementary properties of each polymer.20 These blends can achieve specific mechanical and functional properties such as rigidity and modulus from PHAs, PHBV, and PLA, strength from PLA and PBS, impact strength from PBAT and PCL, and elongation at break from PBS, PBAT, and PCL.21 Furthermore, these polymers also offer unique advantages, such as being renewable resource-based (PHAs, PLA, PBS), having high heat-deflection temperatures (PHAs, PBS, PHB, PHBV), good flowability (PHB, PHBV), and low cost (PLA, PBAT).21
Blending different polymers does have its own drawbacks, including polymer miscibility issues and the need to match the compatibility of polymers to prevent phase separation, which in turn leads to poor performance of the material.22 Despite these challenges, the advantages of biodegradable polymer blends make them an important development in the pursuit of reducing environmental impact while maintaining material performance. The primary objective is to develop biodegradable materials that can effectively replace traditional plastics in a wide range of applications, such as packaging and medical devices, while also maintaining environmental sustainability.23
To address these challenges, recent advancements in biodegradable polymer blends have introduced innovative strategies that enhance material performance and sustainability. A key development is the incorporation of nanocomposites and additives like nanocellulose and nanoclays, which improve mechanical strength, barrier properties, and thermal stability, making these blends more suitable for applications like food packaging and medical devices.24–27 Compatibilizers have also been introduced to enhance the miscibility of different polymers, addressing the issue of phase separation and resulting in materials with better mechanical properties and consistent biodegradation.28,29 Table 1 show a summary of different compatibilizers that are commonly used in polymer blends. Additionally, there is a growing focus on fully bio-based polymer blends, combining renewable resources like PLA and polyhydroxyalkanoates (PHA) with natural materials such as starch, cellulose, and natural fibers. These blends not only lower the carbon footprint but also offer enhanced biodegradability, making them ideal for packaging and agricultural uses. In the biomedical field, blends like PLA and polycaprolactone (PCL) are being tailored for specific therapeutic uses in scaffolds, drug delivery systems and tissue engineering where controlled degradation is crucial.30–32 Moreover, the packaging industry has seen the development of biodegradable blends with improved barrier properties and the ability to interact with packaged products to extend shelf life. These innovations reflect a broader commitment to sustainability in materials science, with ongoing research likely to yield even more sophisticated blends that balance functionality with environmental friendliness.
Table 1 Various compatibilizers, their pros, cons and recommended condition for use
Compatibilizer |
Pros |
Cons |
Recommended use conditions |
Maleic anhydride (MA) |
Improves interfacial adhesion and compatibility46,47 |
Requires careful dosing47 |
PE, PP, and biocomposite; blends with polar fillers46 |
Dicumyl Peroxide (DCP) |
Initiates grafting, enhances crosslinking48,49 |
Over-crosslinking risk; odor and safety issues50 |
Peroxide-initiated grafting49 |
Joncryl® Series |
Chain extension and compatibilization in polyesters51,52 |
|
PLA/PBAT or PET blends; chain-extending, toughening51,52 |
Epoxidized Soybean Oil (ESO) |
Bio-based, improves flexibility, epoxide-reactive55–57 |
Over-plasticization/oil blooming,53 limited thermal stability at high temperature (above 240 °C)54 |
PLA, PBAT, PBS, starch-based flexible blends55,56 |
Triallyltrimesate (TAM) |
Improves crosslinking and stability52,59 |
Can cause loss of Newtonian plateau58 |
Thermoset systems or crosslinked thermoplastics59 |
Tributyrin (TB) |
Plasticizer, improves flexibility60,61 |
Less effective in mechanical performance, reduced mechanical strength and risk of migration60,62 |
PLA blends/composite for packaging or biomedical uses61 |
Polyethylene glycol (PEG) |
Improves hydrophilicity, compatibility in polar systems63–65 |
Migration risk,66 may reduce crystallinity65 |
PEG/PLA, PEG/PCL, PBAT, hydrogels63,64 |
MDI (Diphenylmethane Diisocyanate) |
Reactive; strong coupling/crosslinker for urethanes/polyesters67 |
Toxicity, moisture sensitive68 |
Polyurethane, polyester blends needing strong covalent bonding68 |
TAIC (Triallyl Isocyanurate) |
Crosslinking, thermal stability70,71 |
Can cause brittleness, costly69 |
High-temp or radiation-crosslinked systems70 |
2.1 Biodegradable polymer blends
Generally, polymers are classified as biobased and fossil or petroleum based. Biodegradable polymers can be biobased, and fossil based. Clearly, not all biobased polymers are biodegradable, and not all biodegradable polymers are biobased. The classification of polymers based on origin and biodegradability is presented in Fig. 2, and Table 2 gives the chemical structures and composition of common biodegradable polymers. Biodegradable polymer blends are a promising class of materials with various applications, including in the biomedical and packaging industries. Biodegradable polymer blends are two or more biodegradable polymers mixed to yield a material with tailored properties such as high strength, improved toughness and ductility, controlled degradation rate, and thermal stability without sacrificing their biodegradability. The process of polymer blending can be done with or without a chemical interaction between the individual polymers.33 Biodegradable blends offer a likely new frontier in the search for sustainable materials. They offer special properties, including high performance ability, low costs and easy processability,34 which in turn would broaden their industrial application prospects.35
 |
| Fig. 2 Classification of polymers based on origin and biodegradability. | |
Table 2 Common biodegradable polymers, sources, structures, and molecular formula (structures are drawn in ChemDraw)
Polymer |
Composition |
Structure |
Molecular formula |
PLA |
Lactic acid |
 |
(C3H4O2)n |
PBAT |
Adipic acid, 1,4-butanediol, and terephthalic acid |
 |
(–O–(CH2)4–OCO–(CH2)4–CO–)X[–O–(CH2)4–OCO–C6H4–CO–]y |
PHB |
Hydroxylbutyrate |
 |
(C4H6O2)n |
PHBV |
3-Hydroxybutyrate (3HB) and 3-hydroxyvalerate (3HV) |
 |
(C4H6O2)m (C5H8O2)n |
PHV |
3-Hydroxyvalerate |
 |
C5H10O3 |
PBS |
Butylene succinate |
 |
(C8H12O4)n |
PBSA |
1,4-Butane diol, succinic acid, and adipic acid |
 |
H[O(CH2)4OOC(CH2)2CO]m[O(CH2)4OOC(CH2)4CO] |
PCL |
ε-Caprolactone |
 |
(C6H10O2)n |
PGA |
Glycolic acid |
 |
(C2H4O3) |
Despite the attractive potential of biodegradable polymers, their widespread use is held back by such considerations as costs, mechanical properties, and poor performance.35–37 Developing biodegradable blends with acceptable properties can overcome these problems. However, the choice of individual polymers is dictated by their properties (see Table 3). As an example, a more brittle PLA can be blended with tough and flexible PBAT, to increase elongation at break and flexibility. The main focus of polymer blending is to improve the adhesion between the blends, reduce the interfacial tension, and generate limited inclusion phase size and controlled morphology.38–40 Research has focused on these in recent years which can be effectively addressed by either reactive compatibilization, non-reactive compatibilization and plasticization strategies that can also increase polymer miscibility.41–45
Table 3 Mechanical properties of different biodegradable polymers
Polymer |
Elongation at break (%) |
TS (MPa) |
TM (GPa) |
FS (MPa) |
FM (GPa) |
Notched IS (J m−1) |
Notched IS (kJ m−2) |
Ref. |
TS refers to tensile strength, TM to tensile modulus, FS to flexural strength, FM to flexural modulus, and IS to impact strength. |
PLA |
7.85–10.6 |
40.16–66.02 |
1.85 |
89.2 |
3.8 |
4.04 |
3.19 |
55 and 72 |
PBAT |
385–670 |
21–25.1 |
0.034–0.067 |
4.0–7.5 |
0.09–1.26 |
Non-break |
— |
73–76 |
PBS |
235.7–323 |
26.6–49.5 |
0.29–0.93 |
— |
— |
6.7–22.5 |
6.7 |
77–81 |
PHB |
5–10 |
20–40 |
1.0–3.5 |
29 |
1.0 |
28 |
3 (kJ mm−2) |
82 and 83 |
PBSA |
432.7 |
18.3 |
0.138–0.18 |
19.2 |
0.439 |
∼500 |
32 |
84 and 85 |
PHBV |
2.5–20 |
25–36.6 |
1.2–3.28 |
64.1 |
2.87 |
20.4 |
6 (kJ mm−2) |
74 and 86 |
Bio-PBSA |
313–460.37 |
17.71–23.72 |
0.168–0.281 |
— |
— |
— |
30.7 |
87–89 |
TPS |
60–76.1 |
0.55–5.4 |
0.001–0.077 |
— |
— |
— |
— |
90–92 |
2.2 Key considerations in polymer blend design
One key consideration in designing polymer blends is their miscibility. Miscibility and compatibility in polymer blends are important because they contribute to morphology, properties, and performance.93 One or various properties of different polymers may be desirable in a single polymer blend, or there could be a need to alter the properties of another polymer; here, blending may become a choice. Polymer blends can have homogeneous or heterogeneous properties depending on the individual components in the blend.94,95
A homogeneous polymer blend exhibits uniform properties and composition throughout the material.96 The polymers are miscible at the molecular level, forming a single-phase system with consistent mechanical, thermal, and morphological properties.97 Heterogeneous polymer blends have distinct phases due to the immiscibility of the polymers.98,99 The material's properties can vary depending on the distribution and interaction of these phases, often leading to phase-separated microstructures. Since the main challenge in polymer blends is miscibility, many researchers have reported blends to be partially miscible where the polymers mix to some extent, leading to limited interaction between the components, immiscible where the polymers do not mix, forming a multi-phase system with distinct phase boundaries, or miscible where the polymers mix at the molecular level to form a single-phase system (see Fig. 3). The overall properties of any blend depend on blend composition, the properties of individual components, and the structure and interactions of these components.100 Imre and Pukánszky37 reported that interactions are more important in blends because they determine the mutual solubility of the phases, the thickness, and properties of the interphase formed during blending, as well as the structure of the blend. The glass transition temperature (Tg), the morphologies and solubility parameters such as the Hansen and Hildebrand solubility parameters (see eqn (1) and (2)) of the blends are some of the techniques used to study their miscibility.
|
 | (1) |
 |
| Fig. 3 Types of polymer blend: miscible, partially miscible, and immiscible blends. | |
Ra represents the Hansen solubility parameter, δD, δP, and δH are dispersion component (non-polar van der Waals interactions), polar component (dipole–dipole interactions), and hydrogen bonding components (hydrogen bonding forces).
|
 | (2) |
δ is the Hildebrand solubility parameter, ΔHv, R, and T are heat of vaporisation, gas constant, temperature, and Vm is molar volume.
Other techniques are presented in Table 4. The phase transition and the interfacial tension in the blended polymers can be enhanced through polymer compatibility to promote the dispersion of one phase into another.103
Table 4 Methods for determining miscibility of polymer blends
Method |
Description |
Miscibility indicators |
Ref. |
Scanning electron microscopy (SEM) |
Visualizes the surface morphology of polymer blends |
Absence of phase separation = miscible; distinct domains = immiscible |
55 and 104 |
Transmission electron microscopy (TEM) |
Provides high-resolution images of polymer morphology |
Similar to SEM: uniform morphology = miscible |
74 and 104 |
Hansen solubility parameters (HSP) |
Divides solubility into δD (dispersion), δP (polar), δH (hydrogen bonding) |
Small Ra (Hansen distance) indicates good miscibility (see eqn (1)) |
102 and 105 |
Hildebrand solubility parameter (δ) |
Measures cohesive energy density of polymers |
δ values within 1–2 MPa1/2 suggest miscibility; large differences = immiscibility (see eqn (2)) |
101 |
Differential scanning calorimetry (DSC) |
Measures heat flow during transitions (Tg, Tm, Tc) |
Single Tg = miscible; dual Tg = phase-separated/immiscible |
55, 56, 106–108 |
Atomic force microscopy (AFM) |
Maps topography and phase at nanoscale |
Homogeneous surface = miscible; differences = immiscible |
109–111 |
Optical microscopy |
Light-based visualization of morphology |
Phase boundaries = immiscible; uniform texture = miscible |
112 and 113 |
Theoretical/computational methods |
Uses Flory–Huggins theory to predict blend miscibility via the χ parameter |
Negative or low χ = miscible; high positive χ = immiscible |
105 |
2.3 Blending synthesis methods
Biodegradable polymer blending can be achieved using different methods of synthesis (Table 5). The method chosen will be based on the type of polymer and the applications involved. Melt blending is commonly used where polymers are mixed at high temperatures using twin-screw extruders, single screw extruders and similar equipment.114 The design and architecture of extruders (co-rotating, counter-rotating, screw diameter and length etc.) greatly affect the blend morphology and outcome.115 Process variables such as residence time, temperature, and pressure are also important to achieve desirable properties of the blend; this approach is popular in laboratories and industry because it does not require solvents, can be scaled up and matches extrusion and injection molding processes.116 In some cases, compatibilizers are added to improve polymer compatibility and prevent phase separation.
Table 5 Common methods for synthesizing biodegradable polymer blends
Method |
Mechanism |
Typical polymers |
Pros |
Cons |
Ref. |
Melt blending |
Mixing polymers in molten state |
PLA, PBAT, PHBV, PBS |
Scalable, solvent-free, compatible with extrusion/injection molding |
Needs compatibilizers for immiscible blends |
114–116 |
Solution blending |
Dissolution in common solvent and casting |
PLA, PHBV, PCL |
Good miscibility control, useful for lab studies |
Solvent use, hard to scale up |
117 |
Reactive extrusion (REX) |
Chemical reactions during melt blending using compatibilizers or initiators |
PLA/PBAT, PLA/PHBV |
Enhances miscibility, improves interfacial bonding |
Needs careful control of reaction conditions |
118–120 |
Electrospinning |
Electric field-driven fiber formation from solution or melt |
PLA, PCL, PHAs |
Creates porous nanostructures, suitable for biomedical |
Low throughput, setup complexity |
121 and 122 |
Interpenetrating polymer network (IPN) |
Two interlaced polymer networks, one formed in presence of another |
PLA/PCL, PHAs |
Improved toughness, thermal stability |
Complex formulation, mostly lab-scale |
117 |
Latex blending |
Aqueous dispersion of polymer latex particles then coalesced |
Starch-based blends, PLA emulsions |
Good particle-level mixing, eco-friendly |
Requires drying/coalescence step, limited polymer types |
117 |
Another method of polymer blend synthesis is solution blending. In this method, the polymers are carefully dissolved in common solvent before being cast into films.117 This method is ideal for laboratory-scale applications where researchers can have control over blend uniformity. Solution blending is perceived to be less environmentally friendly because of the solvents used and may be difficult for industrial upscale.
For improving compatibility during blending, reactive extrusion is a valuable method. This process involves in situ chemical reaction while the materials are being processed.118 Materials like maleic anhydride (MA) and Joncryl® chain extender are usually added to improve the compatibility and miscibility of the blend.119,120 Electrospinning is yet another way that polymer blends can be made.121 The process is mainly applied for blends in biomedical applications.122 The polymer solution is spun into thin fibers using an electric field. As a result, very porous structures are obtained, but the method is more complex and limited in throughput.
In addition, interpenetrating polymer networks (IPNs) are a unique approach where two or more polymers are combined in a crosslinked state, with at least one network developing during the crosslinking of the other.117 Because of the interlocked network, IPNs result in mechanically stronger, more thermally stable blends with shape retention which makes them suitable for biomedical and packaging applications. Latex blending is another method of polymer blending where polymers are dispersed as fine particles in an aqueous medium and then coalesced to film.117 Many coatings, adhesives and biodegradable films use latex blending because it allows for better mixing of particles, less solvent use and more flexibility when using fillers or additives.
3. Recent developments in biodegradable polymer blends; mechanical properties, miscibility and compatibility
Many biodegradable polymer blends have been made to widen the scope of application by enhancing their mechanical, thermal, and barrier properties, improving their processability, and tailoring their degradation rates to meet specific requirements. Some biodegradable polymer blends and their properties are presented in Table 6. Biodegradable polymers demonstrate variations in mechanical properties as shown in Table 3, which significantly influence their use in various industries. Polymers such as polylactic acid (PLA) are known for their high tensile strength and stiffness, but brittleness often limits their use in applications requiring flexibility or impact resistance.123 On the other hand, materials like polybutylene adipate terephthalate (PBAT) and polycaprolactone (PCL) exhibit excellent ductility and elongation at break, making them well-suited for flexible films and soft packaging.16 Polybutylene succinate (PBS) offers a more balanced mechanical profile, with moderate strength and flexibility, making it versatile for both rigid and semi-flexible applications.124 Thermoplastic starch (TPS), while attractive due to its low cost and biodegradability, has low mechanical performance and high moisture sensitivity, which often necessitates blending with other polymers to enhance its usability.125 The best performance can be obtained by blending strategies to tailor them for specific end-use applications.
Table 6 Mechanical properties of biodegradable polymers blends
Polymer/bled |
Composition |
Compatibilizer/plasticizer/additive |
Processing method |
Elongation at break (%) |
TS (MPa) |
TM (GPa) |
FS (MPa) |
FM (GPa) |
Notched IS (J m−1) |
Ref. |
TS refers to tensile strength, TM to tensile modulus, FS to flexural strength, FM to flexural modulus, and IS to impact strength. |
PLA/PBS |
80/20 |
— |
Melt blending/compression molding |
14.1 |
60.2 |
1.79 |
— |
— |
8.10 |
77 |
70/30 |
— |
13.2 |
57.5 |
1.55 |
|
|
6.30 |
|
PLA/PBSA |
90/10 |
— |
Extrusion/injection molding |
12.2 |
52.6 |
1.012 |
— |
— |
— |
84 |
|
Joncryl® ADR- |
|
|
|
— |
— |
— |
|
80/20 |
4300 |
121.2 |
41.2 |
0.754 |
— |
— |
— |
|
PLA/bioPBS |
80/20 |
— |
Fused deposition modeling |
— |
64.9 |
3.0 |
— |
— |
— |
134 |
60/40 |
|
57.1 |
2.4 |
|
|
|
|
PLA/PBAT/ESO |
40/60/5 |
Epoxidized soyabean oil |
Injection molding |
84.6 |
30.47 |
— |
64.8 |
3.3 |
24.01 |
55 |
60/40/5 |
|
Injection molding |
75.96 |
19.56 |
— |
39.5 |
2.1 |
13.32 |
|
PLA/PHB |
85/15 |
— |
Extrusion |
140 |
31 |
1.2 |
— |
— |
— |
135 |
PLA/PHB/OLA |
85/15/15 |
Lactic acid oligomer (OLA) |
Extrusion |
35 |
23 |
1.12 |
— |
— |
— |
135 |
85/15/20 |
220 |
18 |
0.95 |
— |
— |
— |
|
85/15/30 |
270 |
19 |
0.59 |
— |
— |
— |
|
PHBV/PBAT |
70/30 |
— |
Injection molding |
3.8 |
27.9 |
2.3 |
51.0 |
2.79 |
28.0 |
74 |
50/50 |
138.6 |
27.5 |
1.6 |
30.6 |
1.38 |
293.9 |
|
30/70 |
345.3 |
26.5 |
6.09 |
15.8 |
5.93 |
no break |
|
PLA/PBAT/Joncryl ADR-4370S |
60/40/0.75 |
Joncryl ADR-4370S |
Reactive compatibilization/compression molding |
579.91 |
40.88 |
— |
— |
— |
29.62 |
136 |
PLA/PBAT/GR (gum rosin) |
80/20/5 parts per hundred (phr.) |
Gum rosin |
Extrusion/Injection molding |
7.3 |
47.3 |
1.44 |
67.2 |
2.5 |
8.3 |
137 |
PLA/TPS/St-PLA (starch grafted poly(lactic acid)) |
30/70/5 |
Starch-graft-poly(lactic acid) |
Compression molding |
4.25 |
10.2 |
0.78 |
— |
— |
— |
138 |
PHBV/PBAT |
70/30/1 |
2,4′-Diphenylmethan-diisocyanate (MDI) |
Twin-screw extrusion/injection molding |
11 |
26 |
1.9 |
|
4.5 |
— |
139 |
Polymer polarity significantly influences the miscibility and functional performance of biodegradable blends.37,126 To better support this discussion, Table 7 presents the solubility parameters of common biodegradable polymers used in these systems. These parameters help predict blend miscibility by quantifying the interaction potential between different polymer phases. The solubility parameter reflects the cohesive energy density of the polymer, and the closer these parameters are between two polymers, the more likely they are to mix well. Polar–polar blends (e.g., PLA with starch or chitosan) typically show good compatibility due to hydrogen bonding and dipole interactions,37 leading to enhanced biodegradability and mechanical properties in applications such as packaging and biomedical devices.
Table 7 Solubility parameters of common biodegradable polymers
Polymer |
Hansen (MPa1/2) |
Hoy (MPa1/2) |
Van Krevelen (MPa1/2) |
Hildebrand (MPa1/2) |
Small's method (MPa1/2) |
Turbidimetric titration (MPa1/2) |
PLA |
20–21.9140 |
21.31141 |
20.66141 |
20.7142 |
— |
— |
|
|
|
|
19.9143 |
|
|
PBAT |
22.1140 |
21.73141 |
21.22141 |
21.9142 |
— |
— |
PHB |
20.7140 |
— |
— |
— |
— |
— |
PHBV |
20.6144 |
21.6145 |
19.9145 |
— |
— |
21.9144 |
PBS |
22.29147 |
— |
— |
— |
— |
22.5146 |
PBSA |
— |
— |
— |
— |
— |
22.3146 |
Bio-PBS |
20.7140 |
— |
— |
— |
— |
— |
Bio-PBSA |
23.3140 |
— |
— |
— |
— |
— |
PGA |
20.55105 |
— |
— |
— |
— |
— |
PCL |
21140 |
— |
— |
— |
— |
— |
TPS |
— |
— |
— |
— |
8.4148 |
— |
Polar–nonpolar blends such as127,128 thermoplastic starch with HDPE or LDPE offer a tunable balance between biodegradability and durability. However, PHB and PHBV have been greatly studied as the most promising biodegradable poly(hydroxy alkanoates). While it is argued that PHB is extremely brittle, PHBV a co-polymer of PHB is highly flexible and can serve to enhance the toughness of PLA.129
3.1 PLA/PHAs-based blends
PLA and PHAs are some of the most studied biodegradable polymers.2,130,131 Their blends have offered significant solutions in many applications. The main aim for blending PLA with PHAs is to reduce its brittleness and enhance toughness by blending flexible and ductile polymers from renewable sources.132 Poly(hydroxyalkanoates) (PHAs) such as PHBV (poly(3-hydroxybutyrate-co-3-hydroxyvalerate)), PHB (poly(hydroxybutyrate), PHO (poly(hydroxy octanoate)), and PHV (poly(hydroxy velerate)) are biodegradable polymers that have received great attention due to their flexibility and toughness.
3.1.1 PLA/PHB. Studies have shown that incorporation of PHB into PLA can significantly increase the blend's thermal stability and enhance properties such as the crystallization rate.132,133 Recently, Gao et al.132 analyzed the performance of PLA/PHB binary blend at 10, 15, 20, 30, and 40 wt% of PHB. They noticed phase separation in the blend that possibly resulted in decreased tensile strength and a linear increase in impact toughness. It was shown that PHB wt% above 20 induced a higher crystallization of the blend, thereby lowering the crystallization temperature. The appearance of two melting peaks confirmed immiscibility of the two polymers.Olejnik et al.149 examined the optimal PLA/PHB ratio for materials with suitable mechanical, processing, and application properties. The PLA/PHB blends with mass ratios of 100/0, 50/10, 50/20, 40/30, 50/50, 30/40, 20/50, 10/50, and 0/100 were prepared using the extrusion process. The result of selected PLA/PHB blends showed two visible melting peaks, which probably signified an immiscible blend or crystalline–amorphous phase. They concluded that the PLA/PHB (50/10) ratio showed optimum performance regarding mechanical and processing properties.
In the processing of PLA/PHB blend, compatibility and miscibility are important factors as confirmed by previous studies. It was shown that the molecular weight plays a significant role in the blending process.150,151 Low molecular weight PLA is miscible with low molecular weight PHB.151 Very importantly, PHB/PLA shows extreme rigidity and brittleness that limit processing and use in flexible applications, thereby requiring the use of plasticizers. In some cases the plasticizers may act to improve the compatibility of the individual polymers.152 Additionally, compatibilizers have been used to impart compatibility and polymer miscibility where needed.
Armentano et al.135 prepared blends of PLA/PHB, using lactic acid oligomer (OLA) as plasticizer. The plasticizer addition was at 15, 20 and 30 wt% by weight, and the blends were extruded for better processability and mechanical properties in flexible film manufacturing. They showed that the lactic acid oligomer plasticizer decreased the PLA/PHB glass transition temperatures, with improved ductile and barrier properties. This is because plasticized blends have higher crystallinity. Field emission scanning electron microscopy (FESEM) images of blend films showed a dispersed PHB phase with a small average diameter in the PLA polymer matrix.
A binary blend of PLA/PHB (60/40) was prepared in the presence of 15 wt% tributyrin (TB) plasticizer.153 The blend without TB was characterized by a rough fracture surface and phase separation, with two microstructure phases which were accompanied by voids. However, the compatibilized blend caused elastic deformation with fuzzy interfaces and no phase separation, suggesting better adhesion in the blend. The addition of TB negatively affected the gas and vapour barrier which increased with plasticization. The mechanical properties showed reduction in the tensile modulus and strength without an improvement in the impact strength. The PLA/PHB blend displayed a double glass transition, with distinct Tg values for PLA and PHB, indicating their immiscibility. Plasticization caused a reduction in the Tg of the blend, indicating enhanced polymer chain mobility. When propylene oxide block copolymer/ethylene oxide (Synperonic (Syn)) and a mix of adjustable lipophilic–hydrophilic balance liquid surfactants (HLB12) were used as compatibilizers in PLA/PHA (70/30) by Anna et al.,154 it was noticed that adding 0.1 wt% Syn reduced PHB particle size and pull-out in the blend, and HLB12 led to a homogeneous morphology with elongated PHB domains. The cold crystallization of PLA was favoured while progressively increasing the blend crystallinity with a noticeable increase in the elastic modulus of the blend.
3.1.2 PLA/PHBV. The properties of PLA/PHBV blends are significantly affected by the conditions of processing and their blending proportion. Typically, the addition of PHBV to PLA improves the toughness and flexibility of the blend while reducing PLA's brittleness. The thermal stability of the blend is also enhanced, making it more suitable for applications requiring higher processing or service temperatures.Typically, the major drawback is the immiscibility of PLA and PHBV, which determines the morphology and properties of the blend. PLA and PHBV are generally considered immiscible due to their different molecular structures and polarities. Polarity is the distribution of electrical charges in polymer molecules. This determines how the molecule interacts with others; this influences their intermolecular forces, such as hydrogen bonding or van der Waals interactions,155 and plays a significant role in polymer miscibility. This immiscibility often leads to phase separation, which can result in a heterogeneous blend with distinct domains of PLA and PHBV. Polymer blends that are immiscible or partially miscible can exhibit phase separation, leading to distinct regions in the blends, and this can influence the crystallization behaviour of the blend.156
It is well known that PLA exhibits a slow rate of crystallization, and this has been addressed by adding nucleating agents.157,158 The crystallization behaviour of PLA can be influenced by the presence of PHBV as it promotes nucleation, leading to an increase in the crystallinity of PLA, which can improve the mechanical properties of the blend. For instance, the mechanical performance limitations of PLA were addressed by blending it with PHBV and subsequent addition of filler to enhance the properties of the resulting melt-blown nonwovens.159 PLA and PHBV were blended in various ratios (10, 20, 30, 40 wt%) and processed with a melt-blown technique. When PHBV was added, the cold crystallization peak of PLA significantly decreased, as PHBV acts as a nucleating agent, promoting crystallization during cooling. However, at PHBV contents higher than 10%, cold crystallization of PLA was almost absent. The blend also experienced a mutual crystallization effect where the crystallization of PLA/PHBV phases was mutually hindered and the enthalpy of crystallization for both phases was lower compared with their pure forms, which was attributed to the immiscibility and phase separation in the blends. Novak et al.160 indicated that the addition of PHBV suppressed the cold crystallization of PLA, which negatively influenced the dimensional stability of injection-molded parts.
In the study by Lo et al.,161 electrospun PLA/PHBV membrane (1
:
1, 3/1) showed partial miscibility with a single wide melting peak, and a shift in the glass transition temperature to lower values was observed in the initial heating cycle. The blends had similar tensile strength and percentage of elongation while the 1
:
1 blend exhibited higher Young's modulus. Also, they showed that the mechanical performance of the PLA/PHBV was controlled by the crystallinity of the blend, as the percentage crystallinity was higher (35.2%) in the 1
:
1 blend compared with the 3
:
1 (13.8%).
The contribution of divergent compatibility in a blend has been shown to affect the scope of application of such blend. Compatibilization is essential for biopolymer blends to become viable alternatives to petroleum-based polymers. Compatibilization improves mixing, stabilizes microstructures, and enhances the synergy between biopolymers by reducing interfacial tension and strengthening adhesion, particularly in the solid state, to enable efficient stress transfer.22 A previous study examined the in situ grafting of 2,5-dimethyl-2,5-di(tert-butylperoxy)hexane, DBPH, a peroxide initiator, via one-step reactive extrusion onto PHBV where the thermal decomposition of PHBV produced unsaturated bonds that graft onto the PLA backbone.161 The optimal 80/20 PLA/PHBV blend with 0.3 wt% DBPH showed significant improvement in mechanical properties by increasing elongation at break to about 15% from about 10% for blend without the initiator while maintaining tensile strength near that of pure PLA. The thermal stability of the blend was improved, which lowered the glass transition temperature, increased crystallization temperatures, and transformed the blend morphology from a sea-island observed without grafting to more layered structure, demonstrating effective toughening and enhanced performance. Similarly, Pouriman et al.162 grafted maleic anhydride (MA) onto PLA and PHBV to enhance their compatibility in blends by using dicumyl peroxide (DCP) as an initiator during melt compounding. MA and DCP improved the interfacial adhesion in the blend and a possibility of chain-shortening reaction was associated with excessive DCP. Morphologically, MA and DCP resulted in finer dispersion of PHBV within the PLA matrix, suggesting improved phase compatibility and reduced domain size. Table 8 shows blends of PLA with polyhydroxyalkanoates with different additives and the findings from the blending process.
Table 8 The processing methods, additive/compatibilizers used, and properties of PHA-based blends
Blend composition |
Aim |
Processing method |
Compatibilizers/plasticizer |
Result summary |
Ref. |
PLA/PHBV (10–40% PHBV) |
Investigates the blending of PLA with PHBV to enhance PLA's foaming behaviour |
Extrusion |
|
• Immiscible blend observed, PHBV exists as dispersed droplets in PLA matrix |
159 |
• The addition of PHBV influenced the crystallization behaviour of PLA, PHBV played nucleating agent, leading to a higher cell nucleation density |
PLA/PHBV 25/75, 50/50, and 75/25 wt% |
Blend property optimization |
Solvent casting method |
|
• PLA/PHBV polymer solutions were immiscible with two noticeable Tg |
163 |
• Mechanical properties decreased with the increasing amount of PHBV. Weak polymer interface of PLA and PHBV |
PLA/PHBV (60 : 40; 50 : 50; and 40 : 60) |
Investigate 3D printability of PHBV/PLA |
Fused filament fabrication (FFF) |
Joncryl grade ADR-4368C |
• Printing parameters depend on viscosity and thermal stability |
164 |
• High temperature of 220 °C and speed of 45 mm s−1 was recommended for decreased viscosity (8–80 Pa s) |
• Joncryl compatibilizer improved interface bonding |
PLA/PHBV (75/25) |
Study the antibacterial effectiveness of PLA/PHBV blend film against L. innocua by adding ferulic, p-coumaric, and protocatechuic antibacterial agents |
Melt-blending and compression molding |
Polyethylene glycol (PEG 1000) |
• 2% of ferulic increases Tg of PLA |
165 |
• Protocatechuic acid caused PHBV supercooling |
• Thermal stability of blend films was observed with ferulic, p-coumaric and protocatechuic acids |
• All three phenolic acids showed antibacterial activity against Listeria innocua and Escherichia coli bacteria |
PLA/PHB |
Investigate better processability and mechanical properties |
Electrospinning |
Lactic acid oligomer (OLA) (10, 15 and 20 wt%) |
• PLA crystallization, decreased viscosity of the blend |
166 |
• 20 wt% OLA produced bead defects and consequent poor mechanical performance |
• The ternary system PLA-PHB with 15 wt% OLA maintained the mechanical properties |
PLA/PHO |
Study the non-isothermal cold crystallization kinetics of PLA/PHO/talc |
Film casting |
|
• The blends displayed two Tg resulting from viscosity mismatch |
51 |
• The blends are not thermodynamically miscible |
• The crystallization of PLA was increased with a higher amount of PHO and talc |
3.1.3 PLA/PHA blend composite. PLA/PHA reinforced composites are innovative biopolymer materials that combine the biodegradability of polylactic acid (PLA) and polyhydroxyalkanoates (PHA) with enhanced mechanical and thermal properties, making them ideal for sustainable applications in various industries. Natural fibers are a good filler material in biodegradable composites. They can add to biodegradability and improve properties such as the mechanical, thermal, and barrier properties of composites. A lot of studies have shown the potential of natural fibers as reinforcement in polymer blends despite their limitations.Boey et al.167 investigated the effect of incorporating spent coffee grounds (SCG) into a poly(hydroxyalkanoate) (PHA) and poly(lactic acid) (PLA) matrix. Various SCG loadings (10–40 wt%) and PHA/PLA ratios were tested for their influence on mechanical properties. A higher SCG content reduced the mechanical performance of the composite in comparison with the blend. The flexural strength dropped from 19 MPa to about 7 MPa. However, adding fungal-treated SCG improved interfacial adhesion, causing the flexural strength to increase by 26%–52% and flexural modulus by 72%–113%. The impact strength was also increased by 56% as a result of better filler–matrix interaction that enhanced stress distribution in the biocomposite.
In PLA/PHBV blends, spent coffee grounds (SCG) were found to influence crystallization behaviour and mechanical properties.160 SCG increased PLA crystallinity from 8% to 17% by acting as a nucleating agent, accelerating crystallization, but reduced crystallization ability in PHBV-rich blends by 10%–12%. This dual effect impacted the mechanical performance, lowering tensile strength by 15%–20% and flexural strength by 1%–10%, depending on the blend composition. Similarly, Marta et al.168 investigated biochar-reinforced composites of PLA/P(3HB-co-4HB) using acetyltributyl citrate (ATBC) plasticizer for sustainable applications. Biochar, used at 10%, 15%, 20%, and 30%, enhanced electrostatic and degradation properties. The surface resistivity decreased significantly from 3.80 × 1012 Ω (no biochar) to 1.32 × 1012 Ω at 30% biochar, improving electrostatic performance. The degradation study revealed that composites with lower biochar content exhibited accelerated degradation, especially in hydrolytic conditions (70 °C), with accelerated PLA loss at biochar content of 20 wt%. However, while SCG accelerated PLA crystallization, biochar incorporation PLA/P(3HB-co-4HB) enhanced stiffness and electrostatic properties but likely impacted flexibility and overall crystallization dynamics, reflecting the complex interplay between fillers, matrix composition, and mechanical performance.
Building on this, Jurczyk et al.169 further investigated PLA/P(3HB-co-4HB)/biochar composites at 0, 10, 15, 20 and 30 wt% using ATBC. Their results showed that the composite became stiffer but weaker and more brittle with increasing amount of biochar. Tensile strength dropped to 27.1 MPa from 35.6 MPa for the pure blend at 30 wt% biochar, a 24% decrease. The tensile modulus doubled from 1230 MPa to 2756 MPa as biochar content increased, while the elongation at break sharply reduced from 12% to 4.2% at 30 wt% biochar, a 65% reduction. The impact resistance decreased by 64% (4.47 kJ m−2 for the neat blend to 1.61 kJ m−2 at 30 wt%). This reduction in strength and impact resistance was attributed to poor bonding between the biochar and polymer matrix, with biochar's porous structure causing cracks to form easily.
3.2 PLA/PBAT-based blends
Blends of PLA and PBAT have been widely researched, with a focus on mechanical properties, rheological behaviour, and biodegradability.106,107,170–173 Blending PLA/PBAT is a good step forward in improving not only the biodegradability of the blend but also its applications. PBAT can significantly increase the flexibility and toughness of the blend but can also lead to reduced tensile strength and modulus. With these in mind, research has been centered around blend compatibility, increasing the miscibility of the blend for better performance. The compatibilized blends of PLA/PBAT can be achieved through physical and reactive compatibilizers. Block or graft copolymers compatible with PLA/PBAT are commonly utilized as physical compatibilizers. Reactive compatibilizers, on the other hand, can chemically link the PLA and PBAT phases, resulting in high interfacial interaction and superior compatibility.174 Wu et al.106 studied the impact of ADR4370S, a polyfunctional epoxy-based chain extender/compatibilizer (0 wt%, 0.3 wt%, 0.6 wt%, 0.9 wt%, and 1.2 wt%) on a blend of PBAT/PLA (95/5) and subsequent supercritical CO2 bead foams. PBAT/PLA compatibility was adjusted by varying the ADR wt%, thus allowing precise foam cell morphology control. The introduction of ADR4370S enhanced the compatibility between PBAT and PLA by facilitating grafting and chain extension reactions, leading to improved interfacial adhesion and a more uniform molecular structure. This resulted in increased viscosity and modulus, contributing to better mechanical strength and elasticity in the blend, which was essential for maintaining foam structure. Differential scanning calorimetry (DSC) analysis showed that the glass transition temperature (Tg) of PBAT increased with higher ADR content as a result of improved compatibility, while crystallinity decreased in the blend because of restricted polymer molecular chain movement. The addition of ADR produced a more homogeneous pore structure, particularly at an optimal ADR content of 0.6 wt%, which achieved the smallest average pore size and the highest foaming ratio.
Epoxidized canola oil (ECO) (5 phr.)-compatibilized PLA/PBAT (80/20, 70/30, and 60/40) blend was prepared by Wahbi et al.107 for application in 3D printing. Their idea was to achieve high impact toughness. In the presence of ECO there were significant improvements in PLA/PBAT compatibility that led to homogeneous PBAT domains within the PLA matrix. The blend containing 30 wt% PBAT showed optimum impact performance of 148 J m−1, an increase of 106% compared with non-compatibilized blends. Analysis of the 3D printed samples showed that there was poor adhesion in the PLA/PBAT strands and larger voids were observed, resulting in brittle failure. The ECO compatibilization induced adhesion of neighboring strands with fewer air voids between deposited strands, resulting in enhanced interlayer adhesion and ductility.
Han et al.108 studied epoxidized soybean oil (ESO)-compatibilized PLA/PBAT blends through reactive compatibilization with varying ESO content at PLA/PBAT ratio of 70/30. The uncompatibilized blend was grossly immiscible with phase separation. The DSC result showed two glass transition temperatures to prove the immiscibility of the blend. They further noted that compatibilization resulted in a single glass transition temperature.
The tensile strength of 70PLA/30PBAT/5ESO increased by 25%, the elongation at break increased more than 6 times, and notched impact strength increased by 400% relative to PLA/PBAT blend.
A biodegradable blend of PLA/PBAT (70/30) was prepared using interfacial stereocomplex crystallites (poly(styrene)-co-glycidylmethacrylate) (i-SCs) having 45% epoxy groups per 100 g by reactive blending.104 The interface compatibilization resulted in enhanced comprehensive performance, which the authors attributed to the “rigid” i-SC, co-continuous morphology. The neat blend showed “sea-island” features having large spherical domain sized-PBAT in the PLA matrix (Fig. 4). This “sea-island” feature had a poor interfacial region, suggesting both polymers were thermodynamically immiscible. Nonetheless, the domain size in the compatibilized blend was greatly reduced due to better interface adhesion (Fig. 5).
 |
| Fig. 4 The morphology of polymer blends using different types of compatibilizers and the effect on polymer phase and miscibility. (1) Field emission scanning electron microscopy (FESEM) images of (a) PLA (b) PLA/PBAT (c) PLA/PBAT5GR, (d) PLA/PBAT10GR, (e) PLA/PBAT15GR, (f) PLA/PBAT20GR, (g) PBAT and (h) PBAT10GR. (2) (a) PLA/PBAT with different reactive compatibilizers; (b) SG-g-PDLA reactive compatibilizer to yield “stable co-continuous PLA/PBAT (DSB) blend and (c) SG-g-PLLA reactive compatibilizer to yield “Sea-island” LSB Blend. (3) Compatibilization effect of ESO on PLA/PBAT.104,108,137 (1) was reproduced from ref. 137 under Creative Commons CC license, copyright 2025. (2) was reproduced from ref. 104 with permission from American Chemical Society, copyright 2025 and (3) was reproduced from ref. 108 with permission from American Chemical Society, copyright 2025. | |
 |
| Fig. 5 The influence of compatibilizer on the morphology of PLA/PBAT blend. SEM and TEM images of 70PLLA/30PBAT blends with and without compatibilizers: (1a1–a3) neat blend, (1b1–b3) (poly(styrene-co-glycidyl methacrylate)-graft-(l-lactide)) (LSB) compatibilized PLA/PBAT blend, and (1c1–c3) (poly(styrene-co-glycidyl methacrylate)-graft-(D-lactide)) compatibilized PLA/PBAT (DSB) blend.104 Figure was reproduced from ref. 104 with permission from American Chemical Society, copyright 2025. | |
Atom transfer radical polymerization via surface-initiation is a grafted-from strategy that can mobilize polymers onto the surface of fillers such as cellulose nanocrystal (CNCs). This method can increase the dispersion of the fillers in matrices and effectively initiate interface compatibility between polymer blends. A dual-purpose incorporation of functionalized cellulose nanocrystal was reported by Sun et al.175 The authors grafted poly(glycidyl methacrylate) (PGMA) (10, 20, and 50%) on the surface of CNC by atom transfer radical polymerization to enhance strength, toughness and compatibility of PLA/PBAT (70/30 wt%) blend. As expected, their result showed that after adding CNC-PGMA, the PLA/PBAT blend (70/30 wt%) reduced in phase size, which they ascribed to improved interfacial compatibility between PLA and PBAT. The PLA/PBAT/CNC1.0-PGMA30 composite exhibited a tensile strength of 49.6 MPa and an elongation at break of 268.5%, which significantly surpassed the mechanical performance of the pure PLA/PBAT and indicated a considerable enhancement in both strength and toughness.
Junwen et al.176 demonstrated that processing parameters and method can significantly impact the properties of a polymer blend. In their work, the blend of PLA/PBAT (80/20) was prepared by combining high mold temperature and strong shear fields during processing. Using the conventional injection molding (CIM) and multi-flow vibration injection molding (MFVIM) techniques led to the formation of highly oriented microstructures, including shish-kebab structures, which increased the blend crystallinity to 48%. The synergistic effect of a high mold temperature and strong shear fields in the MFVIM resulted in tensile strength of 89.07 MPa, tensile modulus of 1664.15 MPa, and impact strength of 10.36 kJ m−2, compared with 53.04 MPa, 1106.47 MPa and 4.61 kJ m−2 for the CIM. Additionally, the Vicat softening temperature (VST) of the blends increased to 149.2 °C, indicating a marked improvement in heat resistance. This synergistic method provides a promising approach for developing high-performance PLA materials, expanding their potential applications as sustainable alternatives to conventional plastics. Table 9 shows blends of PLA with PBAT with different additives and the findings from the blending process.
Table 9 The processing methods, additive/compatibilizers used, and properties of PBAT-based blends
Blend composition |
Aim |
Processing method |
Compatibilizers/plasticizer |
Result summary |
Ref. |
PLA/PBAT (14/86) |
Study antimicrobial properties of polylimonene (PLM) and limonene (LIM) both at 5 wt% and 10 wt% in PLA/PBAT blend for sustainable active food packaging |
Thermo-compression |
|
PLM enhanced antimicrobial and antioxidant activity; 5% PLM preserved visual quality of cherry tomatoes; 10% LIM showed no antifungal activity and promoted fungal growth |
183 |
PLA/PBAT (0, 10, 20, 30, 40 PBAT) |
Study effects of PBAT addition on properties of 3D-printed materials |
Twin-screw extrusion, fused deposition modeling (FDM) |
|
PBAT improved crystallinity and ductility but reduced strength; 30% PBAT had highest elongation (10.15%) |
184 |
PLA/PBAT (70 : 30) |
|
Reactive compatibilization/injection molded |
ESO (epoxidized soybean oil) (0, 0.5, 1, 3, 5, 7, and 9 phr.) |
ESO improved miscibility (single Tg); 5 phr. ESO increased strength by 25%, elongation by 4 folds, impact strength 400% |
108 |
PLA/PBAT (60 : 40, 40 : 60) |
|
Melt blending/injection molding |
ESO (1.5%) |
ESO improved elongation and impact strength; tensile strength decreased; partial miscibility |
53 |
PLA/PBAT (70/30) |
|
Reactive blending |
Interfacial stereocomplex crystallites (poly(styrene)-co-glycidylmethacrylate) (i-SCs) having 45% epoxy groups per 100 g. |
i-SC compatibilization enhanced mechanical properties via co-continuous morphology; uncompatibilized blend was immiscible |
104 |
PLA/PBAT |
|
|
Joncryl ADR-4370S |
ADR improved PBAT dispersion; reduced domain size; unmodified blend showed poor adhesion |
136 |
PLA/PBAT (80/20) |
Study the plasticizing effects of Gum rosin (GR) |
Twin-screw extrusion/by injection molding, film casting |
Gum rosin (GR) (5, 10, 15, and 20 phr) |
GR improved miscibility (Fig. 5), processability and reduced PLA Tm and Tg and increased PBAT domain size |
137 |
While compatibilization has improved the mechanical performance, interface adhesion, and blend morphology of PLA/PBAT blends, it also presents opportunities to fine-tune their degradation properties by optimizing the blend composition and compatibilizer selection. The biodegradability of PLA/PBAT blends has been extensively studied due to their potential as eco-friendly alternatives to traditional plastics, with their formulations designed to balance toughness and degradability for better environmental performance. Research demonstrates that biodegradability depends significantly on environmental conditions and the blend composition.2,177 For instance, under composting conditions, a co-culture of thermophilic bacteria (Pseudomonas G1 and Kocuria G2) degraded PLA/PBAT blends at a rate of 72% within 15 days, primarily through enzymatic activity that targeted ester bonds.178 In soil, higher temperatures (58 °C) enhanced the degradation of the PLA/PBAT blend (8.3 wt% of each), achieving 9.2% and 6.1% degradation for PBAT and PLA.179 However, at a lower temperature of 25 °C, 2.3% of PBAT and 1.7% of PLA degraded respectively, over 33 weeks, while also affecting microbial diversity and soil properties. Similarly, anaerobic digestion (AD) systems showed limited effectiveness for PLA/PBAT blends, with bioplastic bags exhibiting minimal breakdown compared with PLA/PBAT-coated papers, which contributed to biogas production under thermophilic AD conditions.180 Fungal biodegradation of PLA/PBAT blend was also studied using Papiliotrema laurentii S2P4P.181 The fungus simultaneously degraded PLA/PBAT film, decreasing the half-life of PLA/PBAT blends to about 138 days from 3 years, and produced intermediates like adipic acid and lactic acid. Enzymatic degradation of PLA/PBAT blend (20, 40, 60, and 80 wt% PBAT) using cutinase enzyme from Humicola insolens (HiC) demonstrated up to 40% weight loss in PBAT-rich blends within seven days at 70 °C.182 However, the PLA-rich blends show less degradation. Despite these advances, challenges remain, such as incomplete degradation, persistent residues, and environmental variations affecting performance.
3.2.1 PLA/PBAT blend composite. Incorporating turmeric and cinnamon powder as natural fillers in biodegradable PBAT/PLA blend films for use in active packaging showed the potential of cinnamon as an active filler.185 Particularly, cinnamon was effective at reducing UV light transmittance and increasing the surface hydrophobicity of the composite film. Cinnamon contains UV-absorbing compounds like cinnamaldehyde and polyphenols that reduce UV transmittance, while its hydrophobic nature increases surface hydrophobicity. The mechanical properties of the blend improved after adding 5 wt% cinnamon, with elongation increasing by 43%. It was observed that a higher elongation at break could be achieved after 2 reprocessing cycles, which improved filler dispersion in the matrix.Natural cotton stalk (CS) was used as a bio-based filler in PLA/PBAT blends, and the effect on mechanical and barrier properties and compatibility with the blend were studied.186 Adding 2 wt% CS increased tensile strength by 21.5% and elongation by 41.6% (Fig. 8). However, at a higher CS content, filler agglomeration and reduced performance were observed. The CS in PLA/PBAT blend improved the water vapor and oxygen barrier properties of the composite, making it more suitable for packaging applications. Jute natural fiber (JF) in PLA/PBAT blend was reported by Sudha et al.187 to improve mechanical and thermal properties. The Young's modulus of the composite was approximately 1832 MPa with 15%JF, while thermal stability rose from 270 °C in pure PLA to 346.37 °C in the composite (Fig. 8). Impact strength reached 49 J m−1, comparable to ABS-based engineering plastics and outperforming recycled ABS/jute composites. There was improved fiber–matrix bonding, enhanced toughness and crack resistance. Hence, PLA/PBAT/JF could be a potential sustainable alternative to conventional plastics for engineering applications.
The role played by compatibilizers and chain extenders in developing natural fiber-filled biodegradable blend composites is critical in enhancing their mechanical, thermal and barrier properties. Compatibilizers can improve interface adhesion, and chain extenders can increase molecular weight to improve toughness of the composite.
The effect of steam and myristic acid treatment on bamboo fiber (BF) in epoxidized soybean oil (ESO)-compatibilized PLA/PBAT blends was studied by Olonisakin et al.55 Treated BF significantly increased the elongation at break (by up to 644%) and impact toughness (by 370%) of the composites (Fig. 6). There was enhanced fiber dispersion and interfacial bonding with PLA/PBAT and improved thermal stability.
 |
| Fig. 6 Effects of different natural fiber on the mechanical properties of biodegradable blend composite. (1) (a) tensile strength of PLA/PBAT with different jute fiber amounts, (b) tensile stress of PLA/PBAT with different jute fiber amounts. (2) tensile strength and elongation at break of cotton stalk-reinforced PLA/PBAT composite, (3) mechanical properties of bamboo fiber-reinforced PLA/PBAT (elongation at break, tensile strength, flexural strength, flexural modulus and notched Izod impact strength). (1) was reproduced from ref. 187 with permission from Elsevier; copyright 2025, (2) was reproduced from ref. 186, with permission from Willey, copyright 2025, (3) was reproduced from ref. 55, with permission from Elsevier, copyright 2025. | |
Polyethylene grafted maleic anhydride (PE-g-MA) served as coupling agent for better interaction of a PLA/PBAT nanocomposite blend with cellulose nanocrystal.188 A weak interfacial adhesion between the components was observed, due to poor compatibility. The PE-g-MA provided an interface bridge where the MA of PE-g-MA could bond with the –OH of CNC through hydrogen bonding. The CNC and PE-g-MA functioned as nucleating agents, thereby increasing PLA crystallinity and thermal stability.
The addition of rice straw micro-particles to PLA/PBAT blends in the presence of Joncryl® chain extender and maleic anhydride (MA) followed by 500 h accelerated weathering was studied by Mekonnen et al.189 The Joncryl® chain extender was more effective compared with MA in creating a crosslinking network. The addition of rice straw to the blend increased the tensile modulus from 2.3 to 3.4 GPa, a 43% improvement, while the tensile strength decreased (Fig. 6). Rice straw particles served as nucleating agents and lowered the glass transition temperature of the PLA phase from 70 °C to 65 °C. Joncryl®-treated composites retained better mechanical properties after weathering, with a 10–19% reduction in tensile strength compared with 27–30% for MA-treated samples. Similarly, when wood flour and wollastonite was added to PLA/PBAT blends followed by a 1000-hour accelerated weathering (UV exposure), both the blend and the composite exhibited a significant colour change due to the degradation of lignin. The tensile strength and elongation at break declined, with the composite showing a 15% reduction in tensile strength.190
3.3 PBAT/PHA-based blends
Polyhydroxy alkanoates (PHAs) and PBAT are two major biodegradable polymers that have received research attention in the search for eco-friendly materials. PHAs are synthesized by many microorganisms, with biodegradability properties and flexibility, while PBAT is a synthetic aliphatic–aromatic copolyester that exhibits great toughness, flexibility, and processability. Combining these two polymers into biodegradable blends has emerged as a viable approach for improving their overall performance and usability as bioplastics in many applications, without compromising their inherent properties.191
3.3.1 PBAT/PHBV. Biodegradable PHBV/PBAT blends are important in sustainable packaging applications, so the knowledge of their structure–property relationship is important for tailored applications. A study by Resch et al.139 on the blend of PHBV and PBAT utilized biodegradable citrate ester plasticizer (Citrofol BII), and 2,4′-diphenylmethan-diisocyanate (MDI) (1 wt%), as compatibilizer to improve the processing and mechanical properties of the blend. The PHBV/PBAT blend was prepared in ratios of 90/10, 80/20, and 70/30, and processed by injection molding at 180 °C. The results showed that MDI and the citrate ester increased the elongation and toughness of the blend, but with reduced stiffness and strength. Additionally, PBAT increased the viscosity and melt strength of the blend, making it suitable for blown film extrusion. The MDI chemically reacts with the hydroxyl -OH acid –COOH groups of PHBV and PBAT to form urethane and carbamate linkages, thereby improving compatibility and mechanical properties of the blend.A recent study191 explored the blending of biodegradable polyesters, PBAT/PHBV (40/60) using various processing additives like peroxides 2,5-bis(tert-butyl-peroxy)-2,5-dimethylhexane (Luperox 101), 2,5-dimethy-2,5-di(tert-butylperoxy) hexane (Trigonox101) and Joncryl® ADR 4368, an epoxy-based chain extender. The primary aim was to improve the mechanical, thermal, and morphological properties of the blend, addressing issues like immiscibility and poor mechanical performance. The result was a 35% improvement in tensile strength and a 64% improvement in Young's modulus when 0.02 phr. peroxide and 0.3 phr. chain extender were used. These additives significantly increased chain length and formed complex crosslinked network structures that were indicative of non-Newtonian properties in the blend. However, the effectiveness of the chain extender in combination with the peroxides facilitated crosslinking/branching of the polymer chains, resulting in smaller and finely dispersed PBAT droplets.
Jahangiri et al.192 proposed a novel method to enhance the barrier and mechanical performance of PHBV/PBAT/Joncryl blend by coating (dip coating and bar coating) with PHBV. The addition of 0.3 phr. (per hundred resin) Joncryl to the PHBV/PBAT blend improved the blend's miscibility through chemical reactions between the epoxy groups of Joncryl and the hydroxyl/carboxyl end groups of the polymers. This led to the formation of a more homogeneous interfacial layer, leading to an increase in elongation at break by 102%. In terms of barrier properties, PHBV/PBAT/Joncryl bar-coated with PHBV (10–15% coating weight) showed a 48% improvement in water vapor barrier and a 53% improvement in oxygen barrier compared with the uncoated sheet, whereas the dip-coated PHBV/PBAT had little effect on these properties. The elongation at break reduced significantly with the PHBV coating, particularly with the thicker bar-coating that resulted in a more brittle material.
The morphology and performance relationship of PHBV/PBAT biodegradable blends were studied by Zytner et al.74 through injection molding. The PHBV/PBAT blend was varied at 70/30, 50/50, and 30/30 without any compatibilizer. The result showed that 50 wt% of PBAT could improve the brittleness of PHBV with co-continuous configurations in the blend. The SEM results and glass-transition temperature study indicated that the PHBV/PBAT blend had distinct phase separation due to immiscibility (Fig. 7). PHBV crystallinity decreased with the addition of PBAT according to the DSC analysis. This showed that PBAT can hinder the formation of ordered crystals by PHBV. The studied co-continuous morphology could improve mechanical properties by enabling efficient stress distribution, strain transfer, and a balanced contribution of stiffness from PHBV and flexibility from PBAT. These co-continuous configurations could also improve processability by enhancing melt strength and viscosity uniformity, enabling better extrusion and molding performance.193
 |
| Fig. 7 SEM of impact fractured surfaces of PHBV/PBAT with different amounts of each polymer without any compatibilizer. A distinct phase-separated blend can be observed. (i) PHBV/PBAT (70/30), (ii) PHBV/PBAT (50/50), (iii) PHBV/PBAT (30/70), (iv) solvent-etched 70/30PHBV/PBAT, (v) solvent-etched 50/50PHBV/PBAT, (vi) PHBV/PBAT (30/70), after solvent etching.74 Figure was reproduced from ref. 74 under Creative Commons CC license, copyright 2025. | |
Pal et al.194 studied the REX of modified nanoclay-PHBV/PBAT nanocomposite films for use in packaging. Compression molded films were compared with cast film extrusion. First, a 20% nanoclay and PBAT masterbatch was made by melt extrusion. Next, a PHBV/PBAT blend was made using the masterbatch of 0, 0.6, 1.2 and 1.8 wt%, with 40 wt%/60 wt PHBV/PBAT. The PBAT served as a toughener to improve toughness and strain at break of the brittle PHBV. Simultaneously, the nanoclay masterbatch aimed to reinforce the blend and improve PHBV/PBAT interfacial adhesion. According to their findings, the PHBV phase was encircled by the continuous phase PBAT in the PHBV/PBAT due to phase separation. However, the nanoclay improved the dispersion of PHBV in PBAT due to improved interface adhesion. The DSC curves exhibited a distinct crystallization peak for the samples prepared by compression molding and cast film extrusion, showing the formation of uniform crystals. However, the nanoclay induced a subtle hump preceding the crystallization peak, suggesting that the nanoclay promoted imperfect crystals or lamellae in contrast to the PHBV/PBAT blend with uniform crystals. The observed improvement in the crystallization temperature with the addition of nanoclay indicated an impact on the blend's solid-state morphology and crystallization kinetics. This effect suggested enhanced compatibility and interaction between the PHBV and PBAT components, signifying a promising development in the overall thermal behaviour of the nanocomposite material.
3.3.2 PBAT/PHB. Interest in biodegradable blends of PLA/PHAs continues to grow, and the scope of applications is increasing with continuous improvement in research and development. The performance of the blend depends on factors like composition, processing methods, and the presence of additives. Studies have shown that by adjusting the ratio of PBAT to PHB, the properties of the material can be tailored for specific uses. For instance, a higher PBAT content generally enhances flexibility, while a higher PHB content increases stiffness and thermal resistance.In addition to mechanical properties, the biodegradability of PBAT/PHB blends is a key focus. These blends degrade more effectively in natural environments compared with conventional plastics, thanks to the microbial breakdown of PHB and the partial biodegradability of PBAT. Researchers are exploring ways to optimize this biodegradation by studying the role played by soil microbiomes and isolating specific microorganisms that can accelerate the process.
Fernandes et al.195 investigated the biodegradation of PHB/PBAT (45/55) polymer blends in soil and the isolation of novel microorganisms capable of degrading PBAT. The biodegradation was achieved at 27 °C in soil reactors that were inoculated with the soil. At the end of the biodegradation, the microorganisms that degrade PHB/PBAT were isolated. The result revealed that over a six-month incubation period, the bilayer PHB/PBAT film exhibited an average mineralization rate of 47 ± 1%, while the reference material (cellulose) reached 75 ± 1%.
The PHB/PBAT film demonstrated minimal biodegradation within the first 11 days, with less than 2% degradation observed. However, the degradation rate increased until day 52, after which it gradually declined. Additionally, CO2 evolution data indicated higher organic carbon transformation into CO2. Microorganisms such as Streptomyces coelicoflavus, Clonostachys rosea and Aspergillus flavus were found to degrade PHB, while Purpureocillium lilacinum and Aspergillus pseudodeflectus degrade PBAT, respectively, at mesophilic conditions.
Although PBAT and PHB are both biodegradable polymers, their compatibility and miscibility are limited. Generally, PBAT is more flexible and ductile, while PHB is more rigid and brittle. Their miscibility is limited due to differences in their chemical structures and intermolecular interactions.
Costa et al.196 confirmed that PBAT and PHB were immiscible. In their research, PHB, PBAT, and a 1
:
1 blend was prepared, and the non-isothermal melt crystallization kinetics of the samples was determined using DSC and the Avrami, Ozawa, and Mo microkinetic models. They stated that the three models were unable to satisfactorily predict the experimental data for the PBAT/PHB blend. The DSC revealed an immiscible incompatible mix between PHB and PBAT due to their differing degrees of crystallinity, melting temperatures, and peak structures.
Additionally, PBAT/PHB blends with 25 wt%, 50 wt%, and 75 wt% of each component were prepared by Beber et al.197 It was shown that blending PBAT with PHB inhibited the crystallization behaviour of both polymers. A similar phenomenon was also reported by Costa et al.196 DSC results revealed the displacement of the PBAT melting peak by a higher temperature in the PBAT/PHB blends than the neat PBAT.
3.4 PBAT/PHA blend composite
PBAT is a biodegradable aliphatic aromatic polyester having good thermal and mechanical properties, especially impact and elongation at break, and it is a suitable toughening agent in brittle aliphatic polyesters.198 The blends of biodegradable polyhydroxy alkanoantes (PHAs) and PBAT have been researched with immense use in different applications including packaging. To improve the final properties of this blend, researchers in the past decade have incorporated different natural fillers such as lignin,199 babasu fiber,200 hemp,201 starch,202 cellulose nanocrystal,203 etc. Utilizing these materials allowed them to produce materials suitable for a variety of uses while also enabling the industrial and agricultural wastes to have value.199,204
Recently, some research has been documented on natural filler-reinforced PBAT/PHAs. Among this, Yolacan et al.205 did a comprehensive study on PHBV/PBAT, PHBV/PBAT/PHA blends using polyethylene glycol (PEG 400) as plasticizer and adding different amounts of hydroxypropyl methylcellulose (HPMC) (H). Composite films were made through solution casting.
The PHBV/PBAT blend (10/90) exhibited significant mechanical property improvements with a tensile strength of 14 MPa, % elongation of 333%, and elastic modulus of 211 N mm−2, but barrier properties were poor, with an OTR of 297 cc m−2 day−1 and WVTR of 17 g m−2 day−1. The addition of 1% PEG 400, (polyethylene glycol) (P) (PHBV/PBAT/P) improved the tensile strength to 17 MPa, and % elongation increased to 374%, with a slight reduction in elastic modulus (201 N mm−2). The barrier properties (OTR of 233 cc m−2 day−1 and WVTR of 13 g m−2 day−1) were further improved. However, when HPMC at 3% (PHBV/PBAT/P/H3) was added, the maximum tensile strength of 22 MPa, elongation at break of 419%, and elastic modulus of 182 N mm−2 were observed without any significant changes in the barrier properties (OTR of 233 cc m−2 day−1, WVTR of 13 g m−2 day−1).205
The ternary PHBV/PBAT/PHA/P blend demonstrated enhanced mechanical properties, with tensile strength reaching 20 MPa, % elongation of 562%, and elastic modulus of 145 N mm−2, but moderate barrier properties with an OTR of 197 cc m−2 day−1 and WVTR of 9 g m−2 day−1. Increasing HPMC in this blend to 3% (PHBV/PBAT/PHA/P/H3) showed a balance, with tensile strength of 19 MPa, elongation at break at 558%, and an elastic modulus of 187 N mm−2, while barrier properties were excellent (OTR of 197 cc m−2 day−1, WVTR of 9 g m−2 day−1). However, at 5% HPMC (PHBV/PBAT/PHA/P/H5), mechanical performance declined slightly to 17 MPa and 530% elongation at break, though barrier properties remained strong.205
Melendez-Rodriguez et al.206 developed a compostable multilayer film for food packaging, combining polybutylene adipate terephthalate (PBAT) and polyhydroxyalkanoate (PHA) with a barrier layer consisting of CNC (cellulose nanocrystal) and an electrospun hot-tack adhesive made from PHBV. The film exhibited excellent oxygen barrier properties, reducing oxygen permeance by over 90% (from 9.3 × 10–15 to 0.5 × 10–15 m3 m−2 s−1 Pa−1) due to the CNC layer. Water vapor permeance (WVP) was also reduced, with values between 2.0 and 3.6 × 10–12 kg m−2 s−1 Pa−1. While the tensile strength remained stable (around 20 MPa), the CNC significantly reduced the film's flexibility, with elongation at break decreasing by over 90% from 330 and 243% in machine and transverse directions to 27 and 7.6%, respectively. Adhesion between layers was firm, though not very strong, with a peel strength of 0.006 N mm−1. Migration tests showed compliance with food contact standards, and the film fully disintegrated in industrial composting within 60 days, making it a promising sustainable option for high-barrier food packaging.
3.5 Thermoplastic starch-based blends
Extensive research efforts have been directed toward the production of biodegradable starch-based materials for applications ranging from food packaging to potential use in biomedical field. However, the inherent thermodynamic immiscibility resulting from the hydrophobicity of PLA and the hydrophilicity of thermoplastic starch poses challenges, manifesting in poor interfacial adhesion and mechanical properties.207 Research into biodegradable blends of PLA/TPS is continuously evolving, and new applications and processing methods that aid their compatibility are being reported.
The design of biodegradable polymers and TPS blends using molecular dynamics (MD) simulations to understand interfacial properties has revealed the dynamics of blending TPS with PLA, PBS, and PHB.208 Four biodegradable polymers PLA, PBS, PHB, and PBAT were blended with unmodified (nTPS) and citrate-modified (cTPS) TPS. Results showed that PBS, PHB, and PBAT diffused effectively into TPS, forming strong interfaces, while PLA exhibited poor compatibility. PBS and PBAT, particularly when combined with cTPS, demonstrated the highest interfacial fracture energy, indicating robust adhesion due to deep molecular interdiffusion, suppression of voids through electrostatic interactions, and energy absorption through molecular chain conformations. PLA and PHB, with weaker electrostatic interactions, were more prone to interfacial fractures. Diffusion analysis revealed that PBS and PBAT penetrated TPS more effectively, especially with cTPS, whereas PLA showed minimal diffusion.
A novel plant-based compatibilizer—pyrogallic acid (PGA)—was used in PLA/TPS blends with glycerol (4/3/1 wt/wt) as plasticizer via a one-step twin-screw extrusion. The results showed that PGA significantly improved the blend's mechanical properties, including tensile strength, elongation at break, and thermal stability. Specifically, incorporating 1.5 phr. of PGA yielded the highest tensile strength (23.38 MPa) and elongation at break (16.96%), with 24.7% and 233.2% improvements over neat PLA/TPS blends. Additionally, PGA enhanced the blend's crystallinity, water resistance, and interfacial bonding between PLA and TPS due to stable hydrogen bonding and dehydration-induced crosslinking during extrusion.209
The performance of TPS and PBAT blends was enhanced through controlled grafting of a reactive compatibilizer via a two-step blending process, adjusting the grafting density of PBAT onto poly(ethylene glycol methyl ether acrylate-co-glycidyl methacrylate) (EG) compatibilizer (25
:
25
:
1). By varying the pre-grafting time (0, 5, 10, 20, and 40 minutes), the compatibilizer's location within the blend could be controlled. It was shown that the blend with PBAT grafted for 20 minutes demonstrated improved interfacial compatibility, with the smallest TPS domain size and the highest tensile strength (12.2 MPa), storage modulus, and complex viscosity that suggested effective adhesion. Additionally, B20 exhibited superior water barrier properties and the best balance between strength and flexibility.
Similarly, through dynamic vulcanization, the properties of PBAT/TPS biodegradable composite blow molded films were enhanced. Corn starch was plasticized and crosslinked using 1,3,5-tri-2-propenyl-1,3,5-triazine-2,4,6 (1H,3H,5H)-trione (TAIC). The crosslinked TPS was blended with PBAT, improving their compatibility. At optimum content of 2 wt% TAIC and 30 wt% TPS addition, there was significant improvement in tensile strength (25.43 MPa) and elongation at break (580.83%) due to co-crosslinking of TAIC with TPS and TPS/PBAT. Additionally, the blend demonstrated better thermal stability and biodegradability, retaining good processing properties while showing enhanced performance in soil degradation tests.72
The potential application of thermally activated PLA/TPS with enhanced shape memory in biomedicine has been studied by Sessini et al.210 Maleic acid-functionalized oligomeric lactic acid (mOLA) and neat oligomeric lactic acid (OLA) were incorporated as plasticizers (20%). The functionalization of OLA resulted in phase incompatibility and failed to impart the shape memory property to the blend, resulting from the loss of fixed and interchanging phase for shape memory response. However, PLA/TPS with OLA exhibited excellent shape memory properties with a temperature of 45 °C and deformation of 50%. Balancing the plasticizer and compatibilizer is critical for maintaining distinct polymer phases necessary for shape memory function.
Songtipya et al.211 reported on improving the thermo-mechanical properties of thermoplastic TPS and natural rubber (NR) blends by incorporating polyethylene glycol (PEG) and modified natural rubber such as poly(butyl methacrylate)-grafted natural rubber, poly(methyl methacrylate)-grafted natural rubber and epoxidized natural rubber (NR-g-PBMA, NR-g-PMMA, and ENR50N). The impact of different TPS/NR blend ratios, PEG content, and types of modified NR on the blends’ mechanical, thermal, and biodegradation properties were investigated. Results showed that adding modified NR significantly improved the flexibility and toughness of TPS, with the highest toughness observed in the TPS/ENR50 blend (1628 MJ m−3) containing 1.0 wt% PEG. However, excess PEG reduced mechanical properties due to phase separation. Biodegradation tests revealed that the TPS/NR blends biodegraded over 95% within 120 days.
Ávila-Orta et al.212 utilized reactive extrusion (REX) to thermoplastify native cassava starch. They performed in situ chemical modifications by oxidizing the starch with sodium hypochlorite to produce oxidized starch (OS), which was then processed via REX to obtain oxidized thermoplastic starch (O-TS). Additionally, malleated thermoplastic starch (MA-TS) was produced by REX processing of native cassava starch with maleic anhydride (MA) and Luperox 101. A dual-modified thermoplastic starch (O-MA-TS) was created by REX of OS with 2% MA and 0.1% Luperox 101. This process resulted in four types of thermoplastic starch: unmodified thermoplastic starch (TS), oxidized thermoplastic starch (O-TS), malleated thermoplastic starch (MA-TS), and dual-modified thermoplastic starch (O-MA-TS). The obtained TS were blended with PLA at 85% and 65% (w/w) via melt extrusion and subsequently used to produce non-woven fabrics through melt-blowing. According to the SEM findings, the blend with the unmodified TS was immiscible with the PLA, causing large, extremely coarse TS droplets due to phase separation and weak interface bonding. Many other researchers have developed TPS-based blends using different additives that have resulted in interesting findings, as can be seen in Fig. 8.
 |
| Fig. 8 Thermoplastic starch-based blends incorporating different additive and compatibilizers.138,213–216 | |
On the other hand, the dual-modified 65PLA/TS blend and the modified 85PLA/TS blend showed better compatibility, which led to a reduced TS phase with a homogeneous shape. The DTG result suggested improved compatibility and interaction between TS and PLA, especially the dual-modified blends exhibiting a shaper peak in the DTG curve.
3.5.1 TPS-based blend composites. Across different studies, fiber reinforcement was shown to improve mechanical properties, thermal stability, and biodegradability. However, high fiber content sometimes reduced tensile strength, making compatibilizers like maleic anhydride necessary to enhance fiber–matrix interaction.In research on PLA reinforced with rice straw (RS) at 10%, 20%, and 30% weight fractions, the tensile strength decreased as RS content increased, but the addition of maleic anhydride (MA) improved fiber–matrix bonding, resulting in tensile strength increasing by 20% and flexural strength by 14% at 30% RS content. RS fibers increased hydrophilicity, but MA reduced water absorption, improving the material's performance in humid environments.217
For PLA–starch laminates reinforced with RS fibers and active extracts, the incorporation of RS significantly improved tensile and flexural properties.218 These bilayer laminates demonstrated enhanced water vapor and oxygen barrier and fully degraded within 90 days in composting conditions, showing excellent biodegradability, further enhanced by the RS fibers.218 Similarly, for PBAT-based biodegradable films with thermoplastic starch (TPS) and tea polyphenols (TP), films with 30% TPS degraded fully after 180 days, with increased hydrophilicity and water vapor permeability.219 Higher TPS content led to enhanced TP release, improving the films’ antioxidant and antimicrobial properties, though tensile strength decreased with more TPS.
Aouay et al.220 added lignin-containing cellulose nanofibrils (LCNFs) to a blend of film PBAT and TPS (PBAT/TPS) to reduce the water sensitivity of the blend. By adding 10 wt% LCNFs the tensile strength and modulus were improved due to enhanced interaction between the TPS and PBAT phases. There was a reduction in moisture sorption, with up to a 50% decrease in moisture uptake, to achieve an effective reduction in water sensitivity. At a 70/30 PBAT/TPS ratio, adding LCNFs altered the film's morphology. Without LCNFs, TPS dispersed as small, uniform 4–5 μm nodules. As LCNF content increased, TPS particle size grew irregularly, reaching 7, 14, and 20 μm for LCNF contents of 6, 8, and 10 wt%, respectively (Fig. 8).
The addition of 5% and 10% cotton fibers to PLA/TPS composites increased the mechanical properties, with 85% PLA, 10% TPS, and 5% cotton fiber showing the highest water absorption due to increased starch content. Cotton fibers slightly improved the hardness and strength of the composites while maintaining biodegradability.221
The study on mechanical and thermal properties of coir fiber (CF)-reinforced thermoplastic starch (TPS) and poly(butylene adipate-co-terephthalate) (PBAT) composites showed significant improvements in properties with increasing CF content (5%, 10%, 15%, and 20%).222
At 20 wt% CF with alkali treatment, tensile strength improved by 393%, and flexural strength increased by 536% compared with the base blend (Fig. 10). Fiber surface treatment resulted in less pullout and increased fiber–matrix adhesion (Fig. 9). The addition of coir fibers also enhanced thermal stability, as the composites exhibited higher glass transition temperatures and storage modulus, indicating improved thermal performance.222
 |
| Fig. 9 Scanning electron microscopy image showing the influence of different filler and surface treatment on the morphology of TPS-based and PBSA-based blend composites. (a) PBSA-PBAT-WSP, (b) PBSA-PBAT-St, (c) PBSA-PBAT-T, (d) PBSA-PBAT-WSP-T, (e) PBSA-PBAT-St-T, (f) PBSA-PBAT-WSP-St-T, (g) PBSA-PBAT-WSP-T-Compatibilizer, (h) PBSA-PBAT-St-T-compatibilizer, and (i) PBSA-PBAT-WSP-St-T-compatibilizer, (j and k) tensile fracture surface of 20 wt% non-surface-treated coir-fiber reinforced TPS/PBAT, (l and m) tensile fracture surface of 20 wt% alkaline-treated coir fiber-reinforced TPS/PBAT, (n and o) tensile fracture surface of 5 wt% non-surface-treated coir-fiber reinforced TPS/PBAT, (p and q) tensile fracture surface of 5 wt% alkaline-treated coir fiber-reinforced TPS/PBAT composite, cryogenic fractured PBAT-TPS (70–30) containing different lignin-containing cellulose fiber (r) 0 wt%, (s) 6 wt%, (t) 8 wt%, and (u) 10 wt%. Figures a–i were reproduced from ref. 223 and figures j–q were reproduced from ref. 222 with permission from Wiley, copyright 2025. Figures r and s were reproduced from ref. 220 with permission from American Chemical Society, copyright 2025. | |
 |
| Fig. 10 Mechanical properties of PBS, PBSA and bio-PBSA composites. (1) Composites of PBSA/PHBV reinforced with cowpea lignocellulosic fiber (a) tensile strength of injection molded composite, (b) tensile strength of film (c) Young's modulus of injection molded composite, (d) Young's modulus of film composite. (2) PBSA/PHBV with different fillers (a) tensile/flexural modulus, (b) tensile strength/flexural strength, (c) % elongation at break/yield (d) impact strength. (A) PBAT, (B) PBSA-PBAT-WSP, (C) PBSA-PBAT-St, (D) PBSA-PBAT-T, (E) PBAS-PBAT-WSP-T, (F) PBSA-PBAT-St-T, (G) PBSA-PBAT-WSP-St-T, (H) PBSA-PBAT-WSP/T-compatibilizer, (I) PBSA-PBAT-St-Talc-compatibilizer, and (J) PBSA-PBAT-WSP-St-T-compatibilizer. (1) was reproduced from ref. 244 under Creative Commons CC license, copyright 2025 and (2) was reproduced from ref. 223 with permission from Wiley, copyright 2025. | |
3.6 PBS, bio-PBS, bio-PBSA based blends
Polybutylene succinate (PBS) is a biodegradable aliphatic polyester with properties that are comparable to polypropylene. It is produced from succinic acid and 1,4 butanediol.224 PBS can be synthesized through bio-based and petroleum-based pathways. Its biodegradability, biocompatibility, and promising mechanical properties make it suitable for a variety of applications, especially in environmentally friendly packaging and agricultural films. PBS offers good thermal stability, tensile strength, and flexibility compared with other bioplastics. Its biodegradability can be further enhanced when blended with other biopolymers like PLA or TPS.225,226 While PBS is primarily petroleum based, research efforts are increasingly focused on producing bio-based PBS (bio-PBS) from renewable sources such as bio-succinic acid, to enhance its sustainability appeal.
Many studies that incorporate PBS, bio-PBS, and bio-PBSA were driven by the need to develop innovative materials that enhance mechanical performance, improve degradation properties, reduce environmental impact, and create cost-effective solutions in packaging, textiles, and biomedical applications.
One study investigated the effects of blending PBS with PLA and adding an epoxy-functionalized compatibilizer to improve mechanical and barrier properties.227 The addition of organo-montmorillonite Delite®43B (D43B) and random ethylene-methyl acrylate-glycidyl methacrylate terpolymer (ax89) reactive compatibilizer was crucial to improving the oxygen barrier properties of nanocomposite blown films of PBS/PLA. At 3 wt% addition of D43B, the oxygen permeability was reduced by more than 50% and 5 wt% compatibilizer further enhanced oxygen barrier performance and improved elongation at break, in the transverse direction. The blending sequence played a crucial role in determining the compatibility of PBS, PLA, and D43B to influence the mechanical and barrier performance.
The microstructure and mechanical performance of PLA/PBS melt-blown nonwovens has been enhanced by in situ PBS fibrillation technology.228 First a sea-island morphology in PLA/PBS blends was formed due to their immiscibility, where PBS appeared as droplets in the PLA dispersed phase. The PBS droplets were stretched into in situ fibrils under elongational flow during the melt-blown process. The fibrillation process significantly improved the tensile strength and elongation of the blend by 164%, and 672% compared with pure PLA nonwovens. This improvement in mechanical performance, alongside enhanced air and moisture permeability, suggests suitable application of the modified nonwovens in commercial packaging.228 Table 10 lists some recent studies on PBS-based biodegradable polymer blends additives and key findings.
Table 10 PBS-based biodegradable polymer blends
Polymers/additives used |
Processing method |
Key findings |
Ref. |
PLA/PBAT/PBS, peroxide, carbodiimide |
Reactive extrusion with peroxide and carbodiimide, blown film extrusion |
• Reactive extrusion improved rheological properties, for blown film extrusion |
229 |
• Films exhibited tensile strengths up to 57 MPa for tubular films and 53 MPa for champagne films at 30 wt% PBS |
• Seal strength and heat resistance improved with increasing PBS content, suitable for packaging applications involving warm and hot processed meals |
PLA/PBS (70/30 blend) |
Film extrusion, degradation tested in freshwater and seawater |
• PLA degraded most rapidly in warm, light-exposed conditions, generating microplastics |
230 |
• PBS/PLA 70/30 blends showed greater resistance to degradation, without microplastics in cold, dark conditions, indicating slower degradation rates |
PLA/PBS fibers |
Fiber extrusion for textile applications |
• Increasing PBS content enhanced fiber elasticity and mechanical properties |
231 |
• A 7% PBS blend showed a 2.8-fold increase in crystallinity compared with pure PLA |
• Excessive PBS content caused phase separation, reducing mechanical stability |
PHB/bio-based PBS |
Blend preparation for pliable scaffold substrates |
• 50 : 50 PHB/PBS blend showed optimal ductility and hydrophobicity, making it suitable for tissue engineering applications |
232 |
• The blend exhibited enhanced degradation stability and reduced swelling, ideal for biomedical bone tissue scaffold applications |
PBS/PHA |
Marine and compost degradation testing |
• PBS/PHA blends exhibited faster degradation in compost sediment environments (39.5% in 45 days) compared with pure PBS (31.9%) |
233 |
• In marine environments, PBS/PHA degraded by 25.9% and pure PBS by 20.3% in 180 days, making the blend a promising alternative for reducing plastic pollution |
PHBV/PBSA/maleic anhydride-grafted PBSA (PBSA-g-MAH) |
Reactive compatibilization |
• Maleic anhydride grafting improved phase interaction, ductility, toughness, and barrier properties, making the blend suitable for commercial green food packaging applications |
234 |
PHBV/PBSA/epoxy-functionalized chain extender Joncryl® ADR-4468 |
Reactive compatibilization with epoxy-functionalized chain extender (Joncryl) |
• The chain extender improved polymer interaction |
235 |
• Reduced dispersed phase domain size, and increased processability |
• Elongation improved by 45%, transforming the material from brittle to flexible, though impact strength remained unchanged |
PBS/PGA/ESOn-ECD |
Compatibilization using epoxidized soybean oil branched cardanol ether (ESOn-ECD) |
• Adding ESOn-ECD improved compatibility between PBS and PGA, reducing the dispersed phase diameter from 2.74 × 10−14 to 1.97 × 10−14 g cm cm−2 s−1 Pa−1 |
236 |
• Tensile strength increased from 15.3 MPa to 19.4 MPa, and elongation increased from 244.5% to 449.0% |
• Water vapor barrier properties improved significantly |
Blending starch with PBS is a low-cost strategy to stem the high cost of PBS. Rajendran et al.237 argued that PBS is very difficult to market because it is more expensive than fossil-based plastics. Modifying and blending PBS with other polymers could open new applications that reduce unit costs. According to Li et al.238 blending of cassava starch and PBS can be a low-cost PBS approach for developing biocomposites that can replace general-purpose plastics in daily application. For example, a study on modified TPS/PBS/epoxy resin blends suggested that biodegradable modified TPS/PB/epoxy resin blends can be applied in agriculture, packaging and medical fields because of their improved moisture resistance as well as thermal and mechanical properties.239
3.6.1 PBS, bio-PBA, bio-PBSA-based composites. Pei et al.240 investigated four types of lignocellulosic biomass fibers—bagasse, bamboo, rice husk, and rice straw—as filler in a PHB/PBS (7
:
3) matrix to develop bio-composites. The fibers were treated with NaOH to enhance fiber–matrix bonding. The mechanical properties were significantly affected by the type of filler. Bamboo fiber addition to PHB/PBS performed best compared with other filler types, with a bending strength of 19.82 MPa, tensile strength of 12.97 MPa, and impact strength of 4.30 kJ m−2. The bamboo composites were thermally stable, with an initial thermal degradation temperature of 248 °C and a residue of 5.81% after thermal degradation. However, rice straw-based composites performed the poorest, with the lowest mechanical and thermal properties.240Micro fibrillated cellulose (MFC) (0.5%, 0.75%, and 1%) was incorporated into a PLA/PBSA blend for packaging applications. It was found that 0.75% MFC provided the best balance of mechanical, thermal, and barrier properties. The elastic modulus increased as MFC content increased, and the elongation at break peaked at 0.75% MFC, achieving 117% elongation. The addition of MFC also reduced oxygen permeability by 28%, improving the film's barrier properties, although water vapor permeability slightly increased due to the hydrophilic nature of MFC. Cardanol improved MFC dispersion, preventing agglomeration and enhancing film performance. The hot tack strength at 0.75% MFC reached 925 g/15 mm, making it the optimal concentration for achieving a balance of flexibility, strength, and barrier properties, which are critical for sustainable packaging applications.241
The injection-molded blend of biodegradable PBSA/PBAT (60
:
40) polyester blends with WSP (walnut shell powder), St (corn starch), and talc (T) was studied by McNeill et al.223 Compatibilizer such as maleic anhydride grafted PBSA (MA-g-PBSA) was added to improve the compatibility between the different components in the composite.
The addition of 25 wt% talc, WSP, and starch increased the tensile modulus by 234%, 101% and 66% compared with the neat PBSA/PBAT blend (Fig. 10). The flexural modulus also increased by 190%, 97%, and 75%, respectively. These increments were related to the filler type, size, orientation, and shape, with significant contributions to the morphology of the composite. Adding 10 wt% WSP and 15 wt% talc and 5 wt% WSP with 5 wt% St and 15 wt% talc hybrid fillers, the tensile modulus increased by 160% and 162%, while a 147% and 153% increase was observed in the flexural modulus. However, tensile strength decreased. A 25% starch or WSP decreased viscosity and increased water absorption due to the hydrophilic nature of starch.
A dual mechanism involving simultaneous reinforcement with harakeke fiber and compatibilization with dicumyl peroxide (DCP) to improve the mechanical properties of PLA/PBS was proposed by Akindoyo et al.242 The blends without DCP/fiber were observed to have poor miscibility, leading to reduced tensile strength. The addition of harakeke fiber did not significantly improve strength and modulus (Fig. 11). However, the addition of dicumyl peroxide (DCP) as a compatibilizer and harakeke fiber as reinforcement led to a 31% increase in tensile strength and a 148% increase in tensile modulus, with the crystallinity of the blend increasing by 201%. This dual strategy of reinforcement and compatibilization resulted in significantly enhanced mechanical properties compared with unreinforced blends, making it an effective method for improving the performance of biodegradable materials.
 |
| Fig. 11 (1) Tensile strength and tensile modulus of harakeke fiber-reinforced PLA/PBS compatibilized with dicumyl peroxide (DCP), (2) mechanical properties of alkaline-treated coir fiber-reinforced thermoplastic starch/poly(butylene adipate-co-terephthalate) composite (a) tensile strength, (b) Young's modulus, (c) flexural strength, (d) notched Izod impact strength. (1)was reproduced from ref. 242 under Creative Commons CC license copyright 2025 and (2) was reproduced from ref. 222 with permission from Wiley, copyright 2025. | |
Ketata et al.243 hybridized flax fibers (FF) and glass fibers (GF) for reinforcement in a PLA–PBS matrix. The key finding was that hybrid FF/GF resulted in improved mechanical properties compared with using each fiber type alone. The tensile strength of the hybrid composites increased from 42.4 MPa (FF-reinforced composites) to 53 MPa when glass fibers were added. The tensile modulus also improved, increasing from 4.9 GPa for the FF-composite to 5.4 GPa in the hybrid system. However, the addition of glass fibers slightly reduced the elongation at break from 1.7% to 1.5%. Fiber length also played a significant role in reinforcement; while 27% of flax fibers exceeded their critical length in isolation, 34% of fibers surpassed it in the hybrid composite. The glass fibers saw a similar effect, with their percentage increasing from 4% to 19% in the hybrid composite.
In the development of PBS/PLA nanocomposite blown films, Adrar et al.227 added organo-montmorillonite (D43B) and a reactive compatibilizer (ax89), aiming to enhance oxygen barrier properties while maintaining mechanical integrity. Results showed that adding 3 wt% D43B to the PBS/PLA blend reduced oxygen permeability by over 50%, particularly when D43B was localized in the PLA phase. The presence of 5 wt% compatibilizer further improved the oxygen barrier performance. The addition of the compatibilizer ax89 increased elongation at break in the transverse direction, while D43B reduced elongation in the machine direction. The blending sequence significantly influenced the compatibility between PBS, PLA, and D43B, impacting both barrier and mechanical properties.227 Masanabo et al.244 investigated the use of cowpea lignocellulosic fibers as a sustainable filler in bio-composites of PBSA/PHBV for rigid and flexible packaging applications. Composites were prepared by injection molding and film casting. The addition of 10, 20, and 30% of the cowpea fibers increased the tensile strength (17.46 MPa) of injection-molded composites at 30% fiber loading (Fig. 10).
However, for film-cast bio-composites, tensile strength decreased from 18.8 MPa to about 10 MPa due to pore formation. Young's modulus increased with fiber content in both injection-molded and film-cast samples, while the tensile strain decreased as fiber addition restricted polymer chain mobility. The thermal analysis showed a reduction in the onset degradation temperature by up to 30 °C for the 30% fiber-loaded composites.240 Interestingly, water vapor permeability (WVP) and oxygen permeability (OP) were only minimally affected by the fibers, with 20% and 30% fiber bio-composites showing little change compared with fiber-free films.244
4. Role played by inorganic, mineral fillers and nanoparticles in biodegradable blends/composites
There are various types of fillers used in biodegradable composites, each with their unique properties, necessitating their choice of use (see Table 11). Inorganic, mineral, and nanoparticle fillers are widely studied for their ability to enhance stiffness, strength, toughness, flame retardancy, and UV resistance of biodegradable composites, making them suitable for more demanding structural and functional applications.
Table 11 Filler categories used in biodegradable blends/composites244–250
Filler category |
Examples |
Pros |
Cons |
Recommended conditions for use |
Ref. |
Natural fillers |
Turmeric, cinnamon, coffee grounds |
Enhances biodegradability and adds antioxidant properties |
May degrade under high heat; colour may affect aesthetics |
When biodegradability and bioactivity are priorities (e.g., agriculture films) |
160, 167, 185, 248 and 249 |
Rice straw, Flax, Kenaf, Jute, Abaca bagasse, Coir Hemp etc. |
Improves mechanical strength, low cost, biodegradable |
Inconsistent quality due to natural variation |
In applications needing reinforcement and biodegradability |
187, 189, 217, 222, 240, 243 and 255 |
Cellulose (MCC, CNCs, CNFs) |
Reinforces structure, increases thermal stability |
Dispersion issues, may require compatibilizers |
Where structural reinforcement and barrier enhancement are needed (e.g., food packaging) |
175, 241, 256–258 |
Inorganic/mineral fillers |
Talc, calcium carbonate, hydroxyapatite |
Improves stiffness and heat resistance |
Can reduce flexibility; may increase weight |
When stiffness and thermal resistance are needed (e.g., containers) |
250, 251, 252 and 259 |
Silica, montmorillonite |
Improves barrier properties and nucleation effect |
Can lead to agglomeration if not well dispersed |
For high-barrier applications (e.g., sealed food packaging) |
260 and 261 |
Nanoparticles |
Carbon nanotube, graphene oxide, graphene, carbon dots, metal oxides, carbon black |
Enhances mechanical, thermal, and barrier properties at low loadings |
Costly, requires complex surface modification |
For high-performance applications requiring minimal filler loading |
253, 254, 262–264 |
4.1 Inorganic and mineral fillers in biodegradable blends
Significant studies on nanoclay and talc reinforcement in biodegradable composites have been conducted recently aiming to improve their mechanical performance and thermal properties. Nanoclays are efficient for their large surface area and layered structures, providing excellent reinforcement, barrier and mechanical strength to composites. Conversely, talc, a phyllosilicate mineral, is a 2
:
1 tri-octahedral layered silicate complex that consists of an octahedral brucite (Mg(OH)2) layer sandwiched between two tetrahedral silica sheets (Si2O5).245 These minerals have the potential to enhance properties such as barrier and mechanical qualities, making them attractive reinforcing fillers in different industrial applications such as food packaging.
Blend films of biodegradable poly(butylene adipate-co-terephthalate) (PBAT), poly(lactic acid) (PLA), and calcium carbonate (CaCO3) were produced in a one-step melt-blending and film-blowing method. The impact of CaCO3 content and particle size on the properties of these films was evaluated.246 The results showed that increasing the CaCO3 content improved the films’ rheology, tensile performance, and tear resistance. The moisture and oxygen barrier values decreased after incorporating CaCO3 particles. The biodegradable PBAT/PLA films with CaCO3 fillers had excellent mechanical, oxygen, and moisture barrier properties, making them a potential option for food packaging and mulch film in agriculture. Similarly, PLA and inorganic (calcium carbonate (CaCOR3R)) was shown to influence the mechanical, thermal, morphological, and surface properties of PBAT/PLA blend and composites.247 The incorporation of CaCOR3R increased elongation at break compared with PLA, while the blends exhibited a higher elastic modulus, resulting in greater stiffness than the PBAT matrix.
Wang et al.259 evaluated the influence of lignin, calcium carbonate (CaCO3) (PPC), wollastonite (PPW), and talc (PPT) on the properties of PLA/PBAT composites. These fillers improved the mechanical, thermal, and microstructural properties of the composites to varying degrees. The study found that PPC (CaCO3) composites exhibited the best mechanical properties, with tensile strength improved by 15%, impact strength by 27.9%, and bending strength by 32.6%. Thermal analysis revealed that both CaCO3 and wollastonite enhanced thermal stability, with PPC and PPW showing higher crystallinity, particularly for PLA (22.2% for PPC and 24.5% for PPW). Talc, however, decreased thermal stability. These findings suggest that using these fillers, especially CaCO3 and wollastonite, enhances both the mechanical and thermal performance of biodegradable PLA/PBAT composites, making them suitable for various applications.
In the presence of Joncryl chain extender, tensile strength and modulus of PBAT/post-industrial wheat starch (PPWS) increased by 5 and 517%. At the same time, talc functioned as a nucleating material to increase thermal stability of the composites.265
Nanoclay, as earlier mentioned, has great potential as a filler in biodegradable polymer blends. The influence of montmorillonite (MMT) (2 and 5 phr) on the mechanical and thermal properties PBAT/TPS was studied by Peidayesh et al.260 Results showed that adding MMT nanoparticles slightly decreased tensile strength, with tensile stress dropping from 21.3 MPa for the 10% TPS blend to 18.7 MPa with 2% MMT and 18.8 MPa with 5% MMT. However, MMT increased the Young's modulus from 68.7 MPa (no MMT) to 76.0 MPa with 2% MMT, indicating a stiffer material. Elongation at break decreased with MMT addition, confirming the material's increased brittleness. Dynamic mechanical thermal analysis (DMTA) revealed that MMT slightly restricted the mobility of the glycerol-rich regions in the TPS domains, improving stiffness but reducing flexibility.260
4.2 Nanoparticles in biodegradable blends
Nanoparticles are also being applied to reinforce biodegradable polymers, improving their performance for a variety of uses, alongside mineral and inorganic fillers. Nanoparticles are incorporated into polymer blends to improve their properties, especially in the field of food packaging. For example, Khonakdar et al.261 investigated how nanosilica can improve thermal stability and reduce thermo-oxidative degradation of PLA/PBAT blend nanocomposites. The PLA/PBAT blends (90/10 and 75/25 wt/wt) were prepared by adding 1, 3, and 5 phr hydrophilic (HPL) nanosilica and hydrophobic (HPB) nanosilica. The addition of HPB nanosilica improved the thermal stability of nanocomposite. The blend containing 5 phr. HPB nanosilica exhibited the highest degradation activation energy. It was found that nanosilica localized at the PLA/PBAT interface. This enhanced the material's thermal stability. Further observation showed that at higher nanosilica loading, the degradation processes of the composite could be slowed down, making these materials promising for packaging applications.
A multilayer film from PLA/PBAT blends with sodium alginate and coatings of chitosan (CS), zinc oxide nanoparticles (ZnONPs), or silver nanoparticles (AgNPs) for food packaging was assessed for its barrier performance.262 These coatings improved the water vapor and oxygen barrier properties and increased the mechanical strength of the films, especially the AgNPs. The film coated with AgNPs decreased the water vapor transmission rate (WVTR) by 40%, going from 5.77 × 10−14 g cm−2 Pa−1 s−1 for the uncoated film to 3.41 × 10−14 g cm−2 Pa−1 s−1.
In addition, the film exhibited a growth in tensile strength of 8%, going from 264.79 to 286.22 MPa, and displayed strong antibacterial effects on Escherichia coli and Staphylococcus aureus. These films, coated with AgNPs, are ideal for prolonging the shelf life of food products.262
Pozza Junior et al.263 conducted research on using PLA/PBAT/graphite as an electrochemical sensor for detecting the toxic pollutant 2,4,6-trichlorophenol (TCP). The sensors were prepared by electrospinning. It was demonstrated that these sensors had low detection limits (7.84 × 10−4 mol L−1 or 0.0155 mg L−1) and great sensitivity for TCP. Thermal analysis showed that the addition of graphite decreased the crystallinity of the blend and increased the surface area available for analyte adsorption, improving the electrochemical response of the sensor. The sensor had a linear detection range from 1.00 × 10−7 mol L−1 to 2.00 × 10−6 mol L−1 with a correlation coefficient of 0.993, which makes it a good candidate for environmental monitoring.
Zhu et al.266 developed a nanofibrous membrane made of electrospun PHB/PLA composite, incorporating organic photochromic dye (OPD) and silver nanowires (AgNWs), for potential applications in smart textiles, medical care, and counterfeit prevention. The inclusion of 0.05 wt% AgNWs significantly enhanced the antibacterial efficacy against E. coli and S. aureus, exceeding 98%. Breaking stress and breaking strain of the OA-PHB/PLA NFM were 1.61 ± 0.22 MPa and 42.53 ± 5.93%. The membrane had great thermal stability and maintained a filtration efficiency of 99.9% with a minimum pressure drop of 73.4 Pa.
Erick et al.264 sought to enhance the strength, heat resistance, and biodegradability of PLA/PHBV blend by incorporating graphene nanoplatelets (GNP) and multi-wall carbon nanotubes (MWCNT). Adding 2% MWCNT and 5% GNP to the PLA/PHBV blend resulted in a 195% increase in ultimate tensile strength and a 200% increase in strain at break. There was a 30% increase in the elastic modulus of the nanocomposites. PHBV increased thermal degradation temperatures by 10–20 °C. Under composting conditions, the nanocomposites retained good biodegradability, with about 45% degradation in 150 days. Hybrid nanocomposites containing MWCNT and GNP can greatly improve the performance of PLA/PHBV blends while still maintaining biodegradability, making them ideal for sustainable packaging applications.264
Gu et al.267 created m-TiO2 by modifying TiO2 with 3-glycidoxy-propyltrimethoxy-silane and incorporating it into PLLA/PBS blends through reactive blending. The m-TiO2 played the role of interfacial compatibilizer using the epoxy group on its surface. This resulted in simultaneous improvement in the tensile yield strength, notched impact strength and elongation at break. The composite's photocatalytic degradation and antibacterial activity performance were enhanced compared with the neat PLLA/PBS films.
A study on Zinc Oxide Nanoparticle (ZnONP) in PBAT/TPS films with varying ZnONP concentrations (0.5%, 1%, 3%, 5%) and their effects on the films’ properties revealed that higher ZnONP content (5%) significantly improved UV blocking and antimicrobial activity. However, water solubility was reduced, with a decrease to 8.49%. In acidic food simulants, ZnONP migration increased, causing reduced thermal stability, while in aqueous simulants, migration was lower.268
5. Biodegradation of polymer blends and composites
Biodegradable polymer blends and their composites are designed to degrade under natural environmental conditions, and in controlled environments such as industrial and home composting environments, reducing long-term plastic waste accumulation.269 While some polymers degrade well in aerobic conditions like composting, they may not perform as effectively in anaerobic environments such as landfills or anaerobic digestion systems.270 The biodegradability of these polymer blends depends on factors such as their composition, the presence of natural fillers, and external environmental conditions, molecular weight, crosslinking, water solubility, degree of substitution, and crystallinity.271,272 For instance, polylactic acid (PLA) is a widely used biodegradable polymer, but its slow degradation rate in soil and marine environments necessitates blending with other biodegradable materials like thermoplastic starch (TPS) or polybutylene adipate-co-terephthalate (PBAT) to enhance its decomposition.179,273,274 Studies have shown that the addition of natural fibers such as barley straw, rice straw, or nanocellulose further improves biodegradation by increasing water absorption and microbial colonization, which accelerates polymer breakdown.275
In composting environments, PLA/TPS and PBAT-based blends demonstrate significantly enhanced degradation, with some studies reporting over 40% weight loss within 90 days (Table 12). The biodegradability of polymer composites is also influenced by the nature of additives and processing conditions. Plasticizers like polyethylene glycol (PEG) and compatibilizers such as maleic anhydride (MA) can modify the polymer matrix, improving flexibility and water uptake, thereby facilitating microbial degradation. Additionally, factors such as temperature, humidity, and microbial diversity play a crucial role in determining the rate of biodegradation. Research findings indicate that biodegradation at higher temperatures leads to faster polymer breakdown compared with soil burial, where degradation is often limited by environmental fluctuations.
Table 12 Biodegradation studies of polymer blends and composites
Blend composition |
Filler |
Additives |
Biodegradation environment |
Biodegradation conditions |
Standard used |
Findings |
Ref. |
PBAT/PLA (20, 40, 60, 80 wt% PBAT) |
— |
— |
Enzymatic degradation with Humicola insolens cutinase |
70 °C for 7 days |
— |
PBAT-rich blends degraded up to 40% weight loss, while PLA-rich blends showed lower degradation |
182 |
PLA/Tapioca (TS) 65.7/27.9 wt% |
— |
PLA-g-MA, epoxidized palm oil (EPO) |
Soil burial |
23 °C and 30 °C, for 60 days |
ASTM D5988 |
PLA/TS blends exhibited a biodegradation rate of 2.84%. |
276 |
Degradation occurs in the amorphous domains of starch chains |
PLA (80 wt%)/TPS(Cassava) (20 wt%) |
— |
Glycidyl methacrylate (GMA 1 wt%)/benzoyl peroxide (BPO 0.1 wt%) |
Soil burial |
30 ± 2 °C, 90 days |
— |
PLA/TPS showed degradation rate, up to 40% weight loss in 90 days |
277 |
PLA (95 wt%)/PBSA (5 wt%) |
|
Joncryl (3 wt%) |
Soil burial |
30 ± 2 °C, 90 days |
— |
PLA/PBSA film exhibited 8.6% degradation in 90 days of soil burial |
277 |
TPS/natural rubber (90/10 wt%) |
— |
PEG, modified natural rubber |
Soil burial |
120 days |
— |
More than 95% degradation in 120 days; suitable for short-term applications |
211 |
PHB/PBAT (45/55 wt%) |
— |
— |
Soil microbiome |
27 °C, 180 days |
— |
PHB layer degraded faster than PBAT achieving a 47% mineralization in 180 days. PHB increased crystallinity indicated preferential biodegradation of amorphous regions |
195 |
PLA/TPS (56/30 wt%) |
— |
PLA-g-MA (14 wt%) |
Composting |
58 and 37 °C, 180 days |
ASTM D5338-15(2021). |
High mineralization observed at 58 ± 2 °C. |
278 |
At 37 ± 2 °C, improved degradation was observed compared with PLA, with mineralization more than 57% |
TPS/PLA/PBAT (60/30/10 wt%) |
Jute fibers (5, 10, 15 wt%) |
— |
Composting |
37 ± 2 °C and 58 ± 2 °C |
ISO14855-1: 2012 |
Jute fibers acted as a reinforcing agent but slowed down the biodegradation rate |
279 |
Biodegradation rate was faster at 58 ± 2 °C than at 37 ± 2 °C |
PLA/PHB/cellulose paper (75/25 wt%) (sandwich structure) |
Cellulose paper |
— |
Soil burial |
8 months |
— |
16% weight loss under the action of soil microorganisms, water and heat was observed after 8 months of soil burial. |
280 |
PLA/PHB (75/25 wt%) |
— |
— |
Soil burial |
8 months |
— |
PLA/PHB blends achieved a weight loss of 14% over 8 months |
280 |
PLA/PBAT film |
— |
— |
Fungal degradation (Papiliotrema laurentii) in mineral salt medium (MSM) |
30 °C, for 30 days |
— |
Weight loss of 14% observed within 30 days |
181 |
The half-life of film reduced from nearly 3 years to 138 days |
PBAT/TPS (40 to 60 wt% TPS) |
— |
Compatibilizer (MA, PBAT-g-MA) |
Laboratory composting |
58 °C at a 10 mL min−1 air, moisture content less than 50% |
ISO 14855-2:2018 |
PBAT/TPS (40–60 wt%) without compatibilizer, the biodegradation rate was 82–87% after 90days. |
281 |
PBAT/TPS blends with PBAT-g-MA showed reduced biodegradation rates due to the presence of PBAT-g-MA with degradation rate decreasing to 72–74% after 90 days |
PLA/PHB (85/15 wt%) |
Keratin (1–20 wt%) |
Acetyl tributyl citrate (ATBC) (15 wt%) |
Hydrolytic degradation |
0.01 M NaOH solution (pH of 11.8), extra-pure water (MilliQ, pH of 7), 25 °C |
— |
PLA/PHB/Keratin composites showed improved degradation under hydrolytic conditions achieving 50% degradation in 20 days |
282 |
PLA/PHB (85/15 wt%) |
— |
Acetyl tributyl citrate (ATBC) (15 wt%) |
Hydrolytic degradation |
0.01 M NaOH solution (pH of 11.8), extra-pure water (MilliQ, pH of 7), 25 °C |
|
About 50% hydrolysis occurred after 50 days |
282 |
Polymer crystals of PHB were more hydrolytically stable and caused slower degradation |
6. Applications of biodegradable composites
The application of polymer composites depends largely on matrix–reinforcement relationships, the type of polymer matrix, and the type of reinforcement. Nanofillers, synthetic fillers, natural fiber, proteins, starch granules, and biological macromolecules have opened up different applications for reinforced polymer composites. Most recently, and importantly, biodegradable reinforced polymer composites have been utilized across diverse industries due to their versatility (Fig. 12). These composites are gaining increasing attention due to their potential applications in various industries, resulting from their unique properties such as low weight, high strength, environmental friendliness, and biodegradability.
 |
| Fig. 12 Applications of biodegradable polymers and their blends. | |
6.1 Biomedical applications
In medicine, biodegradable reinforced composites are used for implantable devices, such as drug delivery systems and scaffolds for tissue engineering, bones, and dental resin-based composites.130 These materials are attractive due to their biodegradability, which eliminates the need for a second surgery to remove the implant. Their high strength and low weight make them suitable for orthopedic braces and prosthetics. It has been demonstrated that bone regeneration is feasible with a biodegradable piezoelectric poly(L-lactic acid) nanofiber scaffold with ultrasound that can be externally controlled. This is a hybrid of electrical stimulation (ES) and tissue-engineering approaches (biomaterial scaffolds).283 The triblock copolymer of poly(L-lactic acid)-block-aniline pentamer block-poly(L-lactic acid) (PLA-AP) with poly(lactic-co-glycolic acid)/hydroxyapatite (PLGA/HA) scaffold exhibits an improved cell proliferation and better in vitro osteogenesis differentiation for effective bone healing in rabbits.284 In the pursuit of advancing potential biomedical applications, Borah et al.285 developed glycine N-hydroxysuccinimide (NHS) ester-modified polyaniline (PANI)/chitosan nanocomposites for tissue engineering. The nanocomposites showed better fibroblast morphology, proliferation, adhesion, and spreading.
Drug delivery is another area of medicine where biodegradable polymers and their composites are relevant owing to their biocompatibility and degradability. PLA's biodegradability and biocompatibility make it an ideal vehicle for parenteral-controlled drug delivery systems because its microparticles can control the rates of drug delivery that could last for a few days to several weeks and up to a year.286 Several studies have reported the potential efficiency of PLA in active drug delivery, including PEG-PLA loaded with Arenobufagin nanoparticles for enhanced cancer therapy and reduction,287 and PEG-b-PLA micelles and PLGA-b-PEG-b-PLGA sol-gels for drug delivery.288 PLA could present a water-tight coating membrane to significantly reduce the hygroscopicity of mildronate by more than two times without negatively impacting the physical state of the drug, representing a milestone in the cardioprotective drug hygroscopicity, thereby preserving its bioavailability.286,289
6.2 Food packaging
In the packaging industry, reinforced biodegradable matrix composites have been used because they can provide an eco-friendly and sustainable alternative to traditional food packaging materials. These packaging materials are synthesized into film comprising fillers such as starch, cellulose, lignin, and natural fibers, nanoparticles, which micro-organisms in the environment can break down. The reinforced biodegradable matrices are designed to provide superior strength and barrier properties for food packaging, allowing for extended shelf life and improved product safety. Additionally, these matrices are often designed to be compostable, meaning they can be broken down into natural environmental components. These materials have the potential to reduce the amount of plastic waste generated by conventional packaging materials like polyethylene and help to reduce the impact of food packaging on the environment.
PLA is one of the most researched biodegradable polymers for composites. Composite films incorporated with different active fillers have provided functional properties including barrier and antimicrobial functions, and enhanced degradability. Chen et al.290 reported the oxygen permeability of laminated multilayer CNC/PLA film to be 70 times lower than pure PLA film, and the water vapour permeability decreased 7-fold. The CNC was made into a suspension with 15 wt% polyvinyl alcohol (PVA) and then coated on a PLA substrate, followed by lamination. Their idea was that a multilayer composite film could prove successful in improving the barrier of PLA to moisture and oxygen. Also, incorporating cholecalciferol (Vitamin D3) (CC) into PLA solvent-casted composite film could provide an effective barrier to UV-B light at 320 nm, antioxidant activity, and an improved mechanical and oxygen barrier with antibacterial activity against food-borne bacteria (S. aureus and E. coli).291 The functionality of PLA composite film in active packaging has been further improved to include sensing components. Curcumin incorporated into a PLA/poly(propylene carbonate) (PPC) blend has the potential as a smart indicator (sensors that monitor the condition of the food inside the package to provide information to consumers) for food packaging and industrial ammonia (NH3) gas monitoring applications.292 Other fillers in biodegradable polymers have been reported as potential barrier materials in food packaging. These include calcium carbonate, lignin/tannic acid, in PBAT, Pickering emulsions (PE) of essential oils293,294 stabilized by nanocellulose in thermoplastic starch,295 and talc in polybutylene succinate-co-adipate (bio-PBSA).89
6.3 Agriculture
In agriculture and fisheries, composite materials are made into greenhouses for vegetables, flowers, granaries, feed stores, septic tanks, drains, spray, flower pots, milk delivery vehicles, and manure transport vehicles.296 The transition to modern systems of agriculture necessitates the introduction of innovative technologies and environmentally friendly measures to counter the adverse effects of traditional farming. Biodegradable composites have found their use in controlled environment agriculture (CEA). This is the production of plants in an enclosed space like green houses or vertical farms, whereby several environmental variables like temperature, humidity and light among others are carefully controlled.
PLA was found effective as a structural component in macroalgal cultivation (settlement substrates) and cricket rearing (housing) for its resistance, rigidity, and direct material–organism interactions.295
It tolerated corrosive cultivation conditions and provided a suitable substrate with no adverse effect on the macroalgal physiology or nutritional composition, and served as a recyclable shelter for crickets. PHB-reinforced barley fiber composites showed promising potential as a permeable biodegradable composite in agriculture.297 The added fiber increased water uptake capacity to allow water permeation and facilitated biodegradation.
According to Maraveas et al.,298 polysaccharide derivatives (DS), PHB, PCL, PHA, and PLA were shown to be highly biodegradable materials for fabricating anti-insect, anti-hail, and windbreak plastic nets in agriculture. The introduction of biodegradable composites into such systems ensures the optimisation of sustainable production of commodities with a reduction in environmental impact. Such substrates degrade into non-harmful by-products, making controlled environment agriculture more friendly for the environment. This bifunctionality supports the overall purpose of developing sustainable, efficient agricultural systems.
Besides structural applications, biodegradable composites can be used in the development of biodegradable mulches, used for soil management in controlled environment agriculture. Soil moisture, temperature, and weed growth can be effectively controlled using mulching. For instance, TPS, PHA/PLA, and bio-based polybutylene succinate (BioPBS) used as mulching and fruit (tomatoes and peaches) protection bags improved soil quality and reduced blossom end rot.299 The produced peaches had uniform colour (without red blush), a required characteristic for peaches. Their biodegradability in the soil was about 6 months, showing their viability. A problem with traditional plastic mulches is disposal, hence, biodegradable mulches from composite materials provide an effective alternative to traditional non-degradable mulches. During the growing season, these mulches offer the needed agronomic benefits and decompose into residue that enriches the soil humus.
7. Environmental impact assessment
End-of-life considerations and ecotoxicity assessments are crucial aspects in the evaluation of biodegradable polymer blends and composites. Ecotoxicity refers to how chemical, physical, or biological stressors impact ecosystems and organisms including fish, insects, microorganisms, wildlife, and plants.300 As society moves towards more sustainable materials, study of the lifecycle to understand their impact on the environment is essential. It is well understood that biodegradable polymer blends and composites can significantly reduce the burden of traditional plastics on the environment. A thorough assessment of their end-of-life behaviour and potential ecotoxicity is essential. At the end of their useful life, biodegradable polymer blends and composites are designed to undergo industrial or home degradation processes, breaking down into simpler compounds under environmental conditions, such as sunlight, microbial activity, and moisture. The degradation process, rate, and extent are largely dependent on the presence of additives, polymer structure, chemical, molecular weight, and polymer composition. Hence, for precise prediction of the degradation time in different environments, an understanding of the degradation kinetics is essential.
The big question is whether the biodegradation of biodegradable polymer blends and composites raises concerns regarding the degradation of by-products and their potential environmental impact. While biodegradation offers the advantage of reducing the accumulation of plastic waste in landfills and oceans, concerns have been raised about the potential of releasing harmful substances. Under controlled industrial or laboratory conditions, assessing biodegradation based on standard procedures is simpler than in real-life contexts including soils and oceans, which have varying microbial populations.301 According to Ali et al.,302 because biodegradation involves a mixture of identified and unidentified substances that may be hazardous to animals and soil health, it can both pose challenges and offer advantages in complex ecosystems. Unknown substances with roughly 32% cytotoxicity, 23% anti-androgenicity, 42% oxidative stress, and 67% baseline toxicity in bioassays have been shown to be present in bioplastics according to some studies. These compounds have the potential to negatively impact soil and animal health.
Ecotoxicity assessments play a crucial role in evaluating the environmental implications of biodegradable polymer blends and composites. These assessments involve evaluating the potential toxicity resulting from the degradation of various organisms in aquatic and terrestrial environments. Such testing includes a focus on endpoints such as reproductive effects, acute and chronic toxicity, and bioaccumulation potential. To be deemed environmentally benign, these polymers, their blends, and composites must pass toxicity testing and break down into fragments within a given amount of time. However, one of the major setbacks in assessing the ecotoxicity of biodegradable polymers is the lack of standardized testing procedures and regulatory frameworks, unlike the biodegradation and compostability of plastics that have standardized and established testing procedures with clearly outlined regulatory requirements.303 Unlike biodegradation and controlled composting, there seems to be no specific standard for assessing the ecotoxicity of these materials in different environments.
In fact, two ASTM standards, i.e., ASTM D5152-91; practice for water extraction of residual solids from degraded plastics for toxicity testing304 and ASTM D5951-96(2002); standard practice for preparing residual solids obtained after biodegradability standard methods for plastics in solid waste for toxicity and compost quality testing305 were withdrawn in 1998 and 2011, respectively, with no replacements. Hence the assessment of the ecotoxicity of biodegradable materials presents unique complexities as a result of their various compositions and degradation pathways. The multifaceted complexities in the assessment of the ecotoxicity of polymer blends and composites require ecotoxicity testing procedures suited for biodegradable polymers for understanding the possible environmental problems that may be associated with them. A study by Palsikowski et al.306 on ecotoxicological assessment of PLA, PBAT, and their compatibilized blend in soil found that the biodegradable materials and their blends did not exhibit phytotoxic (adverse effects on plant growth, physiology, or metabolism), cytotoxic (quality of being toxic to cells), genotoxic (damaging to DNA or genetic material), or mutagenic (causing genetic mutations) effects on the meristematic cells of Allium cepa, except for a chromosomal aberration index observed in one experiment with the blend 25/75. Sforzini et al.307 used biotests to evaluate the ecotoxicity impact of Mater-Bi, a biodegradable plastic, on soil organisms. The ecotoxicity tests on Mater-Bi biodegradable plastic showed no harmful effects on soil organisms, including autotrophic organisms, Daphnia magna, and Eisenia andrei. In their discussion of the environmental fate and ecotoxicity assessment of biodegradable polymers, Farachi et al.308 emphasized the importance of assessing the possible ecological dangers that may be connected to biodegradable materials. Carteny et al.309 hypothesized that in marine environments, biodegradable microplastics may accumulate more contaminants than conventional microplastics, despite emitting fewer additives. Life cycle analyses (LCAs) are the most complete tools to evaluate the entire lifecycle of recyclable polymers, blends and their composites, starting from the raw material stage to the final disposal stage or end of life. Through LCAs, factors such as greenhouse gas emissions, energy consumption, and ecotoxicity potential are evaluated. These analyses help identify challenges and opportunities to enhance the sustainability of biodegradable materials.
8. Conclusion
8.1 Summary of key findings
In response to the growing environmental awareness and shift towards developing biodegradable materials from renewable sources, research has increasingly focused on addressing the limitations of conventional, non-biodegradable plastics. An important strategy in enhancing the properties and performance of biodegradable polymers is the use of polymer blends and reinforcements, which improve compatibility and miscibility through plasticization, as well as reactive and non-reactive compatibilization. These techniques enhance mechanical strength, flexibility, and overall functionality by reducing interfacial tension in polymer blends, making biodegradable polymer blends more viable alternatives to conventional plastics. For example, PLA/PHAs blends, such as PLA/PHB and PLA/PHBV, exhibit enhanced toughness, flexibility, and thermal stability, making them suitable for packaging, biomedical implants, and 3D printing. PLA/PBAT blends, known for their high flexibility and impact strength, are widely used in food packaging and compostable bags. PBAT/PHA blends, such as PBAT/PHBV, provide improved ductility and biodegradability, making them ideal for sustainable packaging and controlled degradation applications. The incorporation of fillers further enhances these blends, with natural fillers such as coffee grounds, rice straw, and lignocellulosic fibers improving biodegradability and mechanical reinforcement for sustainable composites. Inorganic and mineral fillers, such as nanoclays and calcium carbonate, significantly enhance mechanical strength, barrier properties, and thermal stability, making them suitable for high-performance packaging and structural applications, while nanoparticle reinforcements, such as cellulose nanocrystals (CNCs), graphene, and hydroxyapatite nanoparticles, contribute to improved mechanical, thermal, and antimicrobial properties, particularly in biomedical implants, tissue engineering, and antimicrobial packaging.
9. Future directions
9.1 Integration with circular economy concepts
An innovative method for sustainable manufacturing and waste management is offered by integrating biodegradable polymers and composites into the circular economy (Fig. 13). This paradigm shift is essential for addressing the dual issues of resource depletion and environmental degradation, which are caused by the traditional linear economic model of “take-make-dispose” (Fig. 14). Biodegradable polymers and composites are a great ally of the circular economy, which stresses the ideas of reduce, reuse, repurpose, and recycle.310,311 The circular economy may incorporate biodegradable polymers and composites in many ways, such as waste management, product design and production, and policy formulation. Every stage offers opportunities and challenges to establish a more sustainable cycle of material use and recycling.8
 |
| Fig. 13 Circular economy concept: cradle to cradle for biodegradable polymers, blends, and composites. Figure was made with Biorender. | |
 |
| Fig. 14 Linear economy concept: from cradle to grave. Figure was made with Biorender. | |
The design phase is critical for establishing the lifecycle impact of a product. Designing for biodegradability requires not just choosing appropriate materials, but also considering product usability, durability, and disposal at the end of life.312 Products can be built to decompose under certain conditions, such as in industrial composting facilities, diverting them away from landfills and into usable compost.313 Furthermore, innovation in the combination of biodegradable polymers with natural fibers often results in unique composites with superior properties, expanding their application range and replacing more traditional, non-biodegradable materials in some sectors. The circular economy principle is embodied also in the production process. Efficient, waste-minimizing, and energy-saving manufacturing processes are fundamental steps toward achieving circularity.314,315 Modern manufacturing procedures like additive manufacturing (3D printing) make it possible to design and optimize customized composite systems with no material wastage or energy-demanding machining methods compared with traditional manufacturing. It can reuse recycled materials like plastics, turning waste into new products and supporting a circular economy. This process is energy-efficient, especially for small-batch or complex designs, and reduces the need for global shipping by enabling local, on-demand production. Thus, closed-loop recycling systems including production scrap and end-of-life product recovery, and closed-loop water systems during manufacturing can help to lessen the environmental footprint of these materials.316–318
During utilization, biodegradable polymer composites display an advanced functionality that exhibits variability among sectors like packaging, construction, transport system, and consumer goods.
Applying the circular economy principle, the manufacturers make all types of composite products such as formats and transportation systems durable, repairable and refurbishable, which increases their life cycle. Similarly, the pay-per-use policy such as product-as-a-service models or leasing arrangements which incentivize reuse and resource conservation, encourage the growth of the circular economy mindset in consumers and businesses.8,319 In the circular economy, waste management is essential, and biodegradable polymers, blends, and their composites offer significant advantages.
The two most important methods for handling these materials at the end of their lifespan and turning trash into useful resources like compost and biogas are anaerobic digestion and composting.
The establishment of infrastructure, such as commercial composting facilities, along with customer involvement in appropriate waste sorting, are prerequisites for the success of these initiatives. Moreover, new chemical recycling techniques are being developed that may decompose polymers into their monomers, providing a means of repurposing biodegradable waste to make new materials and thereby closing the loop.
Many studies have focused on the significance of product design, effective utilization, waste management, and reuse. As a step toward a circular economy, sustainable woody-like composites derived from recycled Tetra Pak cellulose and poly(butylene succinate) were fabricated.320 McKeown et al.321 concentrated on PLA chemical recycling, highlighting how end-of-life disposal strategies of biobased polymers are vital for a successful transition to a sustainable circular economy. In the framework of a circular economy, the utilization of waste coffee trash as a renewable source for creating sustainable PBS biocomposites was highlighted.322 The creation of sustainable polymer materials in a way that improves sustainability through material design, waste management, renewable energy sources, biodegradability, biotechnological approaches, and enzymatic recycling within the framework of an international circular (bio)economy can foster efficient resource utilization and management.323
9.2 Biodegradability and recyclability in the circular economy
While biodegradable polymers offer clear environmental benefits especially in mitigating plastic pollution,8,324 the shift toward a circular economy requires a broader, system-level perspective. In this framework, recyclability often takes precedence over biodegradability. This is because materials that can be efficiently reused or reprocessed preserve their resource value and reduce waste generation across multiple life cycles. Blending biodegradable polymers is a common strategy to enhance mechanical, thermal, or barrier properties. However, this approach introduces significant challenges for recycling. Most polymer blends consist of immiscible or only partially miscible phases, leading to material heterogeneity that complicates mechanical or chemical recycling.325 These incompatibilities can result in phase separation, property degradation, and overall reduced recyclability, even when compatibilizers are used. As Ragaert et al.326 and Titone et al.325 clearly emphasized, mechanical recycling remains a key pillar of circular material flow, and mono-material streams are generally more favorable for maintaining quality in closed-loop systems.
Biodegradability, while valuable in open environmental contexts where plastic leakage is inevitable, may not offer the same benefits in structured waste management systems lacking industrial composting infrastructure. In fact, in the hierarchy of circularity, recyclable materials are often prioritized because they better support long-term material retention and reuse.327,328 To align polymer blend development with circular economy goals, future strategies must go beyond traditional property enhancement. This includes designing blends with recyclability in mind through reversible chemistries, dynamic covalent networks, or advanced compatibilization techniques that allow effective reprocessing without compromising performance. Additionally, the concept of ‘design-for-recycling’ should be embedded from the outset, ensuring that material formulations account for their entire lifecycle, not just their immediate utility or degradability.328
Ultimately, achieving true circularity in polymer systems will require striking a careful balance between biodegradability, performance, and recyclability. Rather than viewing these qualities as mutually exclusive, the next generation of materials must be engineered to embody all three to support both environmental sustainability and practical circular economy implementation.
9.3 Potential breakthroughs
The horizon of biodegradable polymer blends and filler-reinforced composites is promising, and some breakthroughs are waiting to happen that could change the field of sustainable materials. With the ongoing efforts by industry and researchers to promote green practices and the emergence of alternatives to petroleum-based plastics, the development and incorporation of biodegradable polymers are expected to be key in making the world greener.
One significant area of future breakthroughs in the development of biodegradable polymer blends is the progress in material design and processing techniques. New developments may lead to biodegradable polymer blends with superior qualities, such as barrier, mechanical and thermal properties, through further refinement of the blend composition, structure, and processing conditions. This would make them more appropriate for a wider range of applications than traditional plastics.
Particularly for biomedical applications, such as implantable devices, drug delivery systems, tissue engineering scaffolds, orthopedic braces and prosthetics, biodegradable reinforced composites are most appropriate. Future progress in the field of biomaterials may be marked by the discovery of new materials with improved biodegradability, biocompatibility, and efficacy for regenerative medicine and personalized healthcare.
The future circular economy integration will be driven by the development of closed-loop systems for biodegradable materials, innovative recycling technologies, and eco-friendly product design approaches that emphasize sustainability at every stage of the product lifecycle.
Author contributions
Kehinde Olonisakin: investigation, methodology, data analysis, visualization, writing—original draft preparation. Amar Mohanty; project conceptualization, investigation, validation, funding acquisition and supervision, writing – Review and editing. Mahendra Thimmanagari: investigation, validation, writing – review and editing. Manjusri Misra: project conceptualization, investigation, methodology, validation, administration, resources, funding acquisition and supervision, writing – review and editing. All authors contributed to the discussion, reviews, and approval of the manuscript for publication.
Conflicts of interest
Authors have no competing-conflicting interests to declare.
Abbreviations
ASTM | American Society for Testing and Materials |
CNCs | Cellulose nanocrystals |
CNPs | Cellulose nanoparticles |
DMA | Dynamic mechanical analysis |
DSC | Differential scanning calorimetry |
ESO | Epoxidized soybean oil |
FESEM | Field emission scanning electron microscopy |
FTIR | Fourier-transform infrared spectroscopy |
GPC | Gel permeation chromatography |
GR | Gum rosin |
HA | Hydroxyapatite |
Hec-g@OA | Hectorite modified with octadecylamine |
LCA | Life cycle analysis |
MA | Maleic anhydride |
MDI | 4,4′-Diphenylmethane diisocyanate |
NC | Nanoclay |
NHS | N-Hydroxysuccinimide |
NMR | Nuclear magnetic resonance |
OLA | Oligomeric lactic acid |
O-MA-TS | Dual modified thermoplastic starch |
PANI | Polyaniline |
PBAT | Polybutylene adipate terephthalate |
PBS | Polybutylene succinate |
PCL | Polycaprolactone |
PE | Polyethylene |
PET | Polyethylene terephthalate |
PGA | Poly(glycoic acid) |
PGMA | Poly(glycidyl methacrylate) |
PGV | Poly(vinyl chloride) |
PHAs | Polyhydroxyalkanoates |
PHB | Poly(hydroxy butyrate) |
PHBHHx | Poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) |
PHBV | Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) |
PHO | Poly(3-hydroxyoctanoate) |
PLA | Polylactic acid |
PLA-AP | Poly(L-lactic acid)-block-aniline pentamer |
PLGA | Poly(lactic-co-glycolic acid) |
PP | Polypropylene |
PPC | Polypropylene carbonate |
REX | Reactive extrusion |
SEM | Scanning electron microscopy |
SI-ATRP | Surface-initiated atom transfer radical polymerization |
TAM | Triallyltrimesate |
TEM | Transmission electron microscopy |
TGA | Thermogravimetric analysis |
TPS | Thermoplastic starch |
UV-Vis | Ultraviolet–visible spectroscopy |
WVTR | Water vapor transmission rate |
XRD | X-ray diffraction |
Data availability
No primary research results, software or code have been included, and no new data were generated or analysed as part of this review.
Acknowledgements
The authors would like to acknowledge financial support from (i) the Ontario Research Fund, Research Excellence Program; Round 11 (ORF-RE 11) from the Ontario Ministry of Colleges and Universities (Project No. 056106 and 056184); (ii) the Ontario Agri-Food Innovation Alliance−Bioeconomy for Industrial Uses Research Program (Project No. 030648, 030671 and 030699); the Ontario Agri-Food Research Initiative (Project No. 056442) (iii) the Natural Sciences and Engineering Research Council (NSERC), Canada Discovery Grants Program (Project No. 401716) and the NSERC, Canada Research Chair (CRC) Program (Project No. 460788); and (iv) the NSERC Alliance Grants Program (Project No. 401769) along with the partner industry Competitive Green Technologies, Lamington, Ontario, Canada (Project No. 055427) in carrying out this study.
References
- V. O. Bulatović, V. Mandić, D. Kučić Grgić and A. Ivančić, Biodegradable Polymer Blends Based on Thermoplastic Starch, J. Polym. Environ., 2021, 29(2), 492–508, DOI:10.1007/s10924-020-01874-w.
- A. Samir, F. H. Ashour, A. A. A. Hakim and M. Bassyouni, Recent advances in biodegradable polymers for sustainable applications, npj Mater. Degrad., 2022, 6(1), 68, DOI:10.1038/s41529-022-00277-7.
- L. K. Ncube, A. U. Ude, E. N. Ogunmuyiwa, R. Zulkifli and I. N. Beas, Environmental Impact of Food Packaging Materials: A Review of Contemporary Development from Conventional Plastics to Polylactic Acid Based Materials, Materials, 2020, 13(21), 4994, DOI:10.3390/ma13214994.
- A. K. Mohanty, F. Wu, R. Mincheva, M. Hakkarainen, J.-M. Raquez, D. F. Mielewski, R. Narayan, A. N. Netravali and M. Misra, Sustainable polymers, Nat. Rev. Methods Primers, 2022, 2(1), 46, DOI:10.1038/s43586-022-00124-8.
- P. European, Plastics – the fast Facts 2023. 2023.
- M. Menossi, M. Misra and A. K. Mohanty, Biodegradable cellulose ester blends: Studies, compatibilization, biodegradable behavior, and applications. A review, Prog. Polym. Sci., 2025, 160, 101919, DOI:10.1016/j.progpolymsci.2024.101919.
- D. Stoica, P. Alexe, A. S. Ivan, D. I. Moraru, C. V. Ungureanu, S. Stanciu and M. Stoica, Biopolymers: Global Carbon Footprint and Climate Change, in Biopolymers: Recent Updates, Challenges and Opportunities, ed. A. K. Nadda, S. Sharma and R. Bhat, Springer International Publishing, Cham, 2022, pp. 35–54. DOI:10.1007/978-3-030-98392-5_3.
- J.-G. Rosenboom, R. Langer and G. Traverso, Bioplastics for a circular economy, Nat. Rev. Mater., 2022, 7(2), 117–137, DOI:10.1038/s41578-021-00407-8.
- J. Cheng, R. Gao, Y. Zhu and Q. Lin, Applications of biodegradable materials in food packaging: A review, Alexandria Eng. J., 2024, 91, 70–83, DOI:10.1016/j.aej.2024.01.080.
- A. R. Ajitha and S. Thomas, Chapter 1 - Introduction: Polymer blends, thermodynamics, miscibility, phase separation, and compatibilization, in Compatibilization of Polymer Blends, ed. A. R. Ajitha and S. Thomas, Elsevier, 2020, pp. 1–29. DOI:10.1016/B978-0-12-816006-0.00001-3.
- Q. Jiao, J. Shen, L. Ye, Y. Li and H. Chen, Poly(oxymethylene)/poly(butylene succinate) blends: Miscibility, crystallization behaviors and mechanical properties, Polymer, 2019, 167, 40–47, DOI:10.1016/j.polymer.2019.01.078.
- H. He, B. Liu, B. Xue and H. Zhang, Study on structure and properties of biodegradable PLA/PBAT/organic-modified MMT nanocomposites, J. Thermoplast. Compos. Mater., 2019, 35(4), 503–520, DOI:10.1177/0892705719890907.
- M. Przybysz-Romatowska, M. Barczewski, S. Mania, A. Tercjak, J. Haponiuk and K. Formela, Morphology, Thermo-Mechanical Properties and Biodegradibility of PCL/PLA Blends Reactively Compatibilized by Different Organic Peroxides, Materials, 2021, 14(15), 4205, DOI:10.3390/ma14154205.
- A. S. Elgharbawy, A.-G. M. El Demerdash, W. A. Sadik, M. A. Kasaby, A. H. Lotfy and A. I. Osman, Enhancing the Biodegradability, Water Solubility, and Thermal Properties of Polyvinyl Alcohol through Natural Polymer Blending: An Approach toward Sustainable Polymer Applications, Polymers, 2024, 16(15), 2141, DOI:10.3390/polym16152141.
- I. Vroman and L. Tighzert, Biodegradable Polymers, Materials, 2009, 2(2), 307–344, DOI:10.3390/ma2020307 . eCollection 2009 Jun.
- S. Shaikh, M. Yaqoob and P. Aggarwal, An overview of biodegradable packaging in food industry, Curr. Res. Food Sci., 2021, 4, 503–520, DOI:10.1016/j.crfs.2021.07.005.
- N. Tripathi, M. Misra and A. K. Mohanty, Durable Polylactic Acid (PLA)-Based Sustainable Engineered Blends and Biocomposites: Recent Developments, Challenges, and Opportunities, ACS Eng. Au, 2021, 1(1), 7–38, DOI:10.1021/acsengineeringau.1c00011.
- L. Henke, N. Zarrinbakhsh, H.-J. Endres, M. Misra and A. K. Mohanty, Biodegradable and Bio-based Green Blends from Carbon Dioxide-Derived Bioplastic and Poly(Butylene Succinate), J. Polym. Environ., 2017, 25(2), 499–509, DOI:10.1007/s10924-016-0828-x.
- N. I. Che Ab Aziz, Y. Zakaria, N. Z. Md Muslim and N. F. Nik Hassan, Emerging and Advanced Technologies in Biodegradable Plastics for Sustainability, in Handbook of Biodegradable Materials, ed. G. A. M. Ali and A. S. H. Makhlouf, Springer International Publishing, Cham, 2023, pp. 533–570. DOI:10.1007/978-3-031-09710-2_21.
- R. Muthuraj, M. Misra and A. K. Mohanty, 5 - Studies on mechanical, thermal, and morphological characteristics of biocomposites from biodegradable polymer blends and natural fibers, in Biocomposites, ed. M. Misra, J. K. Pandey and A. K. Mohanty, Woodhead Publishing, 2015, pp. 93–140. DOI:10.1016/B978-1-78242-373-7.00014-7.
- M. Hassan, A. K. Mohanty and M. Misra, Additive Manufacturing of a Super Toughened Biodegradable Polymer Blend: Structure–Property-Processing Correlation and 3D Printed Prosthetic Part Development, ACS Appl. Polym. Mater., 2024, 6(7), 3849–3863, DOI:10.1021/acsapm.3c03150.
- G. Fredi and A. Dorigato, Compatibilization of biopolymer blends: A review, Adv. Ind. Eng. Polym. Res., 2023, 7(4), 373–404, DOI:10.1016/j.aiepr.2023.11.002.
- A. K. Mohanty, S. Vivekanandhan, J.-M. Pin and M. Misra, Composites from renewable and sustainable resources: Challenges and innovations, Science, 2018, 362(6414), 536–542, DOI:10.1126/science.aat9072.
- A. Anjum, R. Garg, M. Kashif and N. O. Eddy, Nano-scale innovations in packaging: Properties, types, and applications of nanomaterials for the future, Food Chem. Adv., 2023, 3, 100560, DOI:10.1016/j.focha.2023.100560.
- Y. Xu, Z. Wu, A. Li, N. Chen, J. Rao and Q. Zeng, Nanocellulose Composite Films in Food Packaging Materials: A Review, Polymers, 2024, 16(3), 423, DOI:10.3390/polym16030423.
- K. Pal, P. Sarkar, A. Anis, K. Wiszumirska and M. Jarzębski, Polysaccharide-Based Nanocomposites for Food Packaging Applications, Materials, 2021, 14(19), 5549, DOI:10.3390/ma14195549.
- M. Rezghi Rami, S. Forouzandehdel and F. Aalizadeh, Enhancing biodegradable smart food packaging: Fungal-synthesized nanoparticles for stabilizing biopolymers, Heliyon, 2024, 10(18), e37692, DOI:10.1016/j.heliyon.2024.e37692.
- N. Panapitiya, S. Wijenayake, D. Nguyen, C. Karunaweera, Y. Huang, K. Balkus, I. Musselman and J. Ferraris, Compatibilized Immiscible Polymer Blends for Gas Separations, Materials, 2016, 9(8), 643, DOI:10.3390/ma9080643.
- K. Zhan, T. Elder and Y. Peng, Enhancing Polypropylene/Polyethylene Blend Performance Through Compatibilization for A Sustainable Future: A Mini Review Focusing on Establishing Bio-Derived Filler Based Hybrid Compatibilizer System, Macromol. Rapid Commun., 2024, 2400724, DOI:10.1002/marc.202400724.
- A. Afshar, H. Majd, A. Harker and M. Edirisinghe, Tailored binary polymer system PCL-PEO for advanced biomedical applications: Optimization, characterization and in vitro analysis, J. Drug Delivery Sci. Technol., 2024, 95, 105582, DOI:10.1016/j.jddst.2024.105582.
- K. Stafin, P. Śliwa and M. Piątkowski, Towards Polycaprolactone-Based Scaffolds for Alveolar Bone Tissue Engineering: A Biomimetic Approach in a 3D Printing Technique, Int. J. Mol. Sci., 2023, 24(22), 16180, DOI:10.3390/ijms242216180.
- M. Hussain, S. M. Khan, M. Shafiq and N. Abbas, A review on PLA-based biodegradable materials for biomedical applications, Giant, 2024, 18, 100261, DOI:10.1016/j.giant.2024.100261.
- M. Ramesh and M. Muthukrishnan, 25 - Biodegradable polymer blends and composites for food-packaging applications, in Biodegradable Polymers, Blends and Composites, ed. S. Mavinkere Rangappa, et al., Woodhead Publishing, 2022, pp. 693–716. DOI:10.1016/B978-0-12-823791-5.00004-1.
- K. Hamad, M. Kaseem, Y. G. Ko and F. Deri, Biodegradable polymer blends and composites: An overview, Polym. Sci., Ser. A, 2014, 56(6), 812–829, DOI:10.1134/S0965545X14060054.
- L. Rajeshkumar, 19 - Biodegradable polymer blends and composites from renewable resources, in Biodegradable Polymers, Blends and Composites, ed. S. Mavinkere Rangappa, et al., 2022, Woodhead Publishing, pp. 527–549. DOI:10.1016/B978-0-12-823791-5.00015-6.
- S. Mangaraj, A. Yadav, L. M. Bal, S. K. Dash and N. K. Mahanti, Application of Biodegradable Polymers in Food Packaging Industry: A Comprehensive Review, J. Packag. Technol. Res., 2019, 3(1), 77–96, DOI:10.1007/s41783-018-0049-y.
- B. Imre and B. Pukánszky, Compatibilization in bio-based and biodegradable polymer blends, Eur. Polym. J., 2013, 49(6), 1215–1233, DOI:10.1016/j.eurpolymj.2013.01.019.
- V. R. Sastri, Chapter 3 - Materials Used in Medical Devices, in Plastics in Medical Devices, ed. V. R. Sastri, William Andrew Publishing, Boston, 2010, pp. 21–32. DOI:10.1016/B978-0-8155-2027-6.10003-0.
- T. Sekine, K. Fukuda, D. Kumaki and S. Tokito, Enhanced adhesion mechanisms between printed nano-silver electrodes and underlying polymer layers, Nanotechnology, 2015, 26(32), 321001, DOI:10.1088/0957-4484/26/32/321001.
- Z. Chen, C. Cui, C. Jin, X. Li, Y. Zhou, Y. Shao, L. Ma, Y. Zhang and T. Wang, Tough and Thermostable Polybutylene Terephthalate (PBT)/Vitrimer Blend with Enhanced Interfacial Compatibility, Macromol. Rapid Commun., 2023, 44(10), 2200972, DOI:10.1002/marc.202200972.
- R. Muthuraj, M. Misra and A. K. Mohanty, Biodegradable compatibilized polymer blends for packaging applications: A literature review, J. Appl. Polym. Sci., 2018, 135(24), 45726, DOI:10.1002/app.45726.
- B. Wei, Q. Lin, X. Zheng, X. Gu, L. Zhao, J. Li and Y. Li, Reactive splicing compatibilization of immiscible polymer blends: Compatibilizer synthesis in the melt state and compatibilizer architecture effects, Polymer, 2019, 185, 121952, DOI:10.1016/j.polymer.2019.121952.
- G. H. Fredrickson, S. Xie, J. Edmund, M. L. Le, D. Sun, D. J. Grzetic, D. L. Vigil, K. T. Delaney, M. L. Chabinyc and R. A. Segalman, Ionic Compatibilization of Polymers, ACS Polym. Au, 2022, 2(5), 299–312, DOI:10.1021/acspolymersau.2c00026.
- M. H. Nguyen, Compatibilization of rubber/polyethylene blends. 2008 Search PubMed.
- J. Andrzejewski, J. Cheng, A. Anstey, A. K. Mohanty and M. Misra, Development of Toughened Blends of Poly(lactic acid) and Poly(butylene adipate-co-terephthalate) for 3D Printing Applications: Compatibilization Methods and Material Performance Evaluation, ACS Sustainable Chem. Eng., 2020, 8(17), 6576–6589, DOI:10.1021/acssuschemeng.9b04925.
- W. Nonkrathok, T. Trongsatitkul and N. Suppakarn, Role of Maleic Anhydride-Grafted Poly(lactic acid) in Improving Shape Memory Properties of Thermoresponsive Poly(ethylene glycol) and Poly(lactic acid) Blends, Polymers, 2022, 14(18), 3923, DOI:10.3390/polym14183923.
- C. Kaynak and Y. Meyva, Use of maleic anhydride compatibilization to improve toughness and other properties of polylactide blended with thermoplastic elastomers, Polym. Adv. Technol., 2014, 25(12), 1622–1632, DOI:10.1002/pat.3415.
- M. K. Dhal, A. Banerjee, K. Madhu, A. B. Gole, A. Kumar and V. Katiyar, Effect of dicumyl peroxide-coated sawdust in melt-processed polylactic acid/polycaprolactone biocomposite for enhanced thermomechanical properties, Int. J. Biol. Macromol., 2025, 308, 142404, DOI:10.1016/j.ijbiomac.2025.142404.
- C. G. Xu, X. G. Luo, X. R. Zhuo and L. L. Liang, Research on Crosslinking of Polylactide Using Low Concentration of Dicumyl Peroxide, Mater. Sci. Forum, 2009, 620–622, 189–192, DOI:10.4028/www.scientific.net/MSF.620-622.189.
- A. Thitithammawong, A. Thitithammawong, K. Sahakaro and J. Noordermeer, Multifunctional peroxide as alternative crosslink agents for dynamically vulcanized epoxidized natural rubber/polypropylene blends, J. Appl. Polym. Sci., 2008, 111, 819–825, DOI:10.1002/APP.29129.
- A. Altınbay, C. Özsaltık, D. Jahani and M. Nofar, Reactivity of Joncryl chain extender in PLA/PBAT blends: Effects of processing temperature and PBAT aging on blend performance, Int. J. Biol. Macromol., 2025, 303, 140703, DOI:10.1016/j.ijbiomac.2025.140703.
- Y. Kahraman, B. Özdemir, B. E. Gümüş and M. Nofar, Morphological, rheological, and mechanical properties of PLA/TPU/nanoclay blends compatibilized with epoxy–based Joncryl chain extender, Colloid Polym. Sci., 2023, 301(1), 51–62, DOI:10.1007/s00396-022-05043-4.
- B. Sun, B. I. Chaudhary, C.-Y. Shen, D. Mao, D.-M. Yuan, G.-C. Dai, B. Li and J. M. Cogen, Thermal stability of epoxidized soybean oil and its absorption and migration in poly(vinylchloride), Polym. Eng. Sci., 2013, 53(8), 1645–1656, DOI:10.1002/pen.23417.
- R. Turco, S. Mallardo, D. Zannini, A. Moeini, M. D. Serio, R. Tesser, P. Cerruti and G. Santagata, Dual role of epoxidized soybean oil (ESO) as plasticizer and chain extender for biodegradable polybutylene succinate (PBS) formulations, Giant, 2024, 20, 100328, DOI:10.1016/j.giant.2024.100328.
- K. Olonisakin, H. Lin, P. Haojin, W. Aishi, H. Wang, R. Li, Z. Xin-Xiang and W. Yang, Fiber treatment impact on toughness and interfacial bonding in epoxidized soya bean oil compatibilized PLA/PBAT bamboo fiber composites, Mater. Today Commun., 2024, 38, 107790, DOI:10.1016/j.mtcomm.2023.107790.
- R. Turco, S. Mallardo, D. Zannini, A. Moeini, M. Serio, R. Tesser, P. Cerruti and G. Santagata, Dual role of epoxidized soybean oil (ESO) as plasticizer and chain extender for biodegradable polybutylene succinate (PBS) formulations, Giant, 2024, 20, 100328, DOI:10.1016/j.giant.2024.100328.
- X. Huang, X. Yang, H. Liu, S. Shang, Z. Cai and K. Wu, Bio-based thermosetting epoxy foams from epoxidized soybean oil and rosin with enhanced properties, Ind. Crops Prod., 2019, 139, 111540, DOI:10.1016/J.INDCROP.2019.111540.
- P. Tiwary, N. Najafi and M. Kontopoulou, Advances in peroxide-initiated graft modification of thermoplastic biopolyesters by reactive extrusion, Can. J. Chem. Eng., 2021, 99(9), 1870–1884, DOI:10.1002/cjce.24080.
- H. Simmons and M. Kontopoulou, Hydrolytic degradation of branched PLA produced by reactive extrusion, Polym. Degrad. Stab., 2018, 158, 228–237, DOI:10.1016/j.polymdegradstab.2018.11.006.
- M. L. Iglesias-Montes, M. Soccio, F. Luzi, D. Puglia, M. Gazzano, N. Lotti, L. B. Manfredi and V. P. Cyras, Evaluation of the Factors Affecting the Disintegration under a Composting Process of Poly(lactic acid)/Poly(3-hydroxybutyrate) (PLA/PHB) Blends, Polymers, 2021, 13(18), 3171, DOI:10.3390/polym13183171.
- N. Petchwattana, P. Kerdsap and B. Sukkaneewat, Plasticization of Poly(Vinyl Chloride) by Non-Carcinogenic Bio-Plasticizers, Key Eng. Mater., 2020, 862, 99–103, DOI:10.4028/www.scientific.net/KEM.862.99.
- E. Kassegn, B. Sirhabizu, T. Berhanu, B. Buffel and F. Desplentere, A study of the mechanical, thermal and rheological properties of sisal fiber-reinforced polylactic acid bio-composites with tributyl 2-acetylcitrate as a plasticizer, J. Thermoplast. Compos. Mater., 2024, 37(11), 3516–3539, DOI:10.1177/08927057241235649.
- H.-Y. Yu, C. Wang and S. Y. H. Abdalkarim, Cellulose nanocrystals/polyethylene glycol as bifunctional reinforcing/compatibilizing agents in poly(lactic acid) nanofibers
for controlling long-term in vitro drug release, Cellulose, 2017, 24(10), 4461–4477, DOI:10.1007/s10570-017-1431-6.
- Z. Xu, X. Qiao, K. Sun, Y. Chen and H. Liu, Biodegradable compatibilizer modified corn stover/poly(butylene adipate-co-terephthalate) composites, Polym. Compos., 2024, 45(5), 4650–4661, DOI:10.1002/pc.28087.
- O. S. Aremu, L. Katata-Seru, Z. Mkhize, T. L. Botha and V. Wepener, Polyethylene glycol (5,000) succinate conjugate of lopinavir and its associated toxicity using Danio rerio as a model organism, Sci. Rep., 2020, 10(1), 11789, DOI:10.1038/s41598-020-68666-z.
- A. Rezaei Kolahchi, A. Ajji and P. J. Carreau, Enhancing hydrophilicity of polyethylene terephthalate surface through melt blending, Polym. Eng. Sci., 2015, 55(2), 349–358, DOI:10.1002/pen.23910.
- H. Jiang, Z. Zheng, W. Song and X. Wang, Moisture-cured polyurethane/polysiloxane copolymers: Effects of the structure of polyester diol and NCO/OH ratio, J. Appl. Polym. Sci., 2008, 108(6), 3644–3651, DOI:10.1002/app.27343.
- A. V. Wisnewski, J. Liu and C. A. Redlich, Antigenic changes in human albumin caused by reactivity with the occupational allergen diphenylmethane diisocyanate, Anal. Biochem., 2010, 400(2), 251–258, DOI:10.1016/j.ab.2010.01.037.
- H. Zhao, J. Chen, H. Zhang, Y. Shang, X. Wang, B. Han and Z. Li, Theoretical study on the reaction of triallyl isocyanurate in the UV radiation cross-linking of polyethylene, RSC Adv., 2017, 7(59), 37095–37104, 10.1039/C7RA05535H.
- A. Matsumoto, T. Kubo, K. Watanabe, H. Aota, Y. Takayama, A. Kameyama and T. Nakanishi, Difference in temperature effect on the polymerizations between triallyl isocyanurate and its isomer triallyl cyanurate, Eur. Polym. J., 2000, 36(4), 673–677, DOI:10.1016/S0014-3057(99)00129-9.
- A. Matsumoto, S. Ogawa, T. Matsuda, A. Ueda, H. Aota, T. Fujii and H. Toridome, Further Discussion on Correlation between Brittleness and Inhomogeneous Network Structure of Cross-Linked Resins Originating in Specific Polymerization Behavior of Triallyl Isocyanurate, Macromolecules, 2008, 41(21), 7938–7945, DOI:10.1021/ma800974s.
- K. Cai, X. Wang, C. Yu, J. Zhang, S. Tu and J. Feng, Enhancing the Mechanical Properties of PBAT/Thermoplastic Starch (TPS) Biodegradable Composite Films through a Dynamic Vulcanization Process, ACS Sustainable Chem. Eng., 2024, 12(4), 1573–1583, DOI:10.1021/acssuschemeng.3c06847.
- J. Jian, Z. Xiangbin and H. Xianbo, An overview on synthesis, properties and applications of poly(butylene-adipate-co-terephthalate)–PBAT, Adv. Ind. Eng. Polym. Res., 2020, 3(1), 19–26, DOI:10.1016/j.aiepr.2020.01.001.
- P. Zytner, A. K. Pal, F. Wu, A. Rodriguez-Uribe, A. K. Mohanty and M. Misra, Morphology and Performance Relationship Studies on Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)/Poly(butylene adipate-co-terephthalate)-Based Biodegradable Blends, ACS Omega, 2023, 8(2), 1946–1956, DOI:10.1021/acsomega.2c04770.
- Y. Liu, S. Liu, Z. Liu, Y. Lei, S. Jiang, K. Zhang, W. Yan, J. Qin, M. He, S. Qin and J. Yu, Enhanced mechanical and biodegradable properties of PBAT/lignin composites via silane grafting and reactive extrusion, Composites, Part B, 2021, 220, 108980, DOI:10.1016/j.compositesb.2021.108980.
- M. Menossi, M. Misra and A. K. Mohanty, Effect of Peroxide Compounds on Biodegradable Blends Based on Poly(butylene adipate-co-terephthalate)/Plasticized Cellulose Acetate, ACS Appl. Polym. Mater., 2024, 6(17), 10202–10217, DOI:10.1021/acsapm.4c01226.
- T. Zhao, J. Yu, X. Zhang, W. Han, S. Zhang, H. Pan, Q. Zhang, X. Yu, J. Bian and H. Zhang, Thermal, crystallization, and mechanical properties of polylactic acid (PLA)/poly(butylene succinate) (PBS) blends, Polym. Bull., 2024, 81(3), 2481–2504, DOI:10.1007/s00289-023-04848-9.
- O. Valerio, M. Misra and A. K. Mohanty, Statistical design of sustainable thermoplastic blends of poly(glycerol succinate-co-maleate) (PGSMA), poly(lactic acid) (PLA) and poly(butylene succinate) (PBS), Polym. Test., 2018, 65, 420–428, DOI:10.1016/j.polymertesting.2017.12.018.
- S. Wu, Y. Zhang, J. Han, Z. Xie, J. Xu and B. Guo, Copolymerization with Polyether Segments Improves the Mechanical Properties of Biodegradable Polyesters, ACS Omega, 2017, 2, 2639–2648, DOI:10.1021/acsomega.7b00517.
- A. Boonprasertpoh, D. Pentrakoon and J. Junkasem, Effect of PBAT on physical, morphological, and mechanical properties of PBS/PBAT foam, Cell. Polym., 2019, 39(1), 31–41, DOI:10.1177/0262489319873859.
- Y. Yan and Q. Dou, Effect of Peroxide on Compatibility, Microstructure, Rheology, Crystallization, and Mechanical Properties of PBS/Waxy Starch Composites, Starch/Staerke, 2021, 73(3–4), 2000184, DOI:10.1002/star.202000184.
- M. K. M. Smith, D. M. Paleri, M. Abdelwahab, D. F. Mielewski, M. Misra and A. K. Mohanty, Sustainable composites from poly(3-hydroxybutyrate) (PHB) bioplastic and agave natural fibre, Green Chem., 2020, 22(12), 3906–3916, 10.1039/D0GC00365D.
- B. McAdam, M. Brennan Fournet, P. McDonald and M. Mojicevic, Production of Polyhydroxybutyrate (PHB) and Factors Impacting Its Chemical and Mechanical Characteristics, Polymers, 2020, 12(12), 2908, DOI:10.3390/polym12122908.
- D. Lascano, L. Quiles-Carrillo, R. Balart, T. Boronat and N. Montanes, Toughened Poly (Lactic Acid)—PLA Formulations by Binary Blends with Poly(Butylene Succinate-co-Adipate)—PBSA and Their Shape Memory Behaviour, Materials, 2019, 12(4), 622, DOI:10.3390/ma12040622.
- N. Harder, A. Rodriguez-Uribe, M. R. Snowdon, M. Misra and A. K. Mohanty, Hop natural fiber-reinforced poly(butylene succinate-co-butylene adipate) (PBSA) biodegradable plastics: effect of fiber length on the performance of biocomposites, Mater. Adv., 2023, 4(6), 1502–1514, 10.1039/D2MA00831A.
- A. El-Hadi, R. Schnabel, E. Straube, G. Müller and S. Henning, Correlation between degree of crystallinity, morphology, glass temperature, mechanical properties and biodegradation of poly (3-hydroxyalkanoate) PHAs and their blends, Polym. Test., 2002, 21(6), 665–674, DOI:10.1016/S0142-9418(01)00142-8.
- R. Tejada Oliveros, J. Martinez, D. Sanoguera, N. Montanes and L. Carrillo, Development and characterization of high environmental performance composites of Bio-PBSA and short hemp fibers from different compatibilization strategies, J. Compos. Mater., 2022, 1(1), 1–5, DOI:10.23967/r.matcomp.2022.
- C. Dolza, E. Gonga, E. Fages, R. Tejada-Oliveros, R. Balart and L. Quiles-Carrillo, Green Composites from Partially Bio-Based Poly(butylene succinate-co-adipate)-PBSA and Short Hemp Fibers with Itaconic Acid-Derived Compatibilizers and Plasticizers, Polymers, 2022, 14(10), 1968, DOI:10.3390/polym14101968.
- D. Nath, A. K. Pal, M. Misra and A. K. Mohanty, Biodegradable Blown Film Composites from Bioplastic and Talc: Effect of Uniaxial Stretching on Mechanical and Barrier Properties, Macromol. Mater. Eng., 2023, 308(12), 2300214, DOI:10.1002/mame.202300214.
- D. S. Lai, S. A. Adnan, A. F. Osman, I. Ibrahim and H. Haq, Mechanical Properties of Thermoplastic Starch Biocomposite Films with Hybrid Fillers, J. Phys.:Conf. Ser., 2021, 2080(1), 012011, DOI:10.1088/1742-6596/2080/1/012011.
- A. de S. M. de Freitas, J. S. Rodrigues, C. C. Maciel, A. A. F. Pires, A. P. Lemes, M. Ferreira and V. R. Botaro, Improvements in thermal and mechanical properties of composites based on thermoplastic starch and Kraft Lignin, Int. J. Biol. Macromol., 2021, 184, 863–873, DOI:10.1016/j.ijbiomac.2021.06.153.
- C. Caicedo and H. L. Pulgarin, Study of the Physical and Mechanical Properties of Thermoplastic Starch/Poly(Lactic Acid) Blends Modified with Acid Agents, Processes, 2021, 9(4), 578, DOI:10.3390/pr9040578.
- C. C. Sarath, R. A. Shanks and S. Thomas, Chapter 1 - Polymer Blends, in Nanostructured Polymer Blends, ed. S. Thomas, R. Shanks and S. Chandrasekharakurup, William Andrew Publishing, Oxford, 2014, pp. 1–14. DOI:10.1016/B978-1-4557-3159-6.00001-8.
- I. Khan, M. Mansha and M. A. J. Mazumder, Polymer Blends, in Functional Polymers, ed. M. A. Jafar Mazumder, H. Sheardown and A. Al-Ahmed, Springer International Publishing, Cham, 2018, pp. 1–38. DOI:10.1007/978-3-319-92067-2_16-1.
- T. d. Naylor, 20 - Permeation Properties, in Comprehensive Polymer Science and Supplements, ed. G. Allen and J. C. Bevington, Pergamon, Amsterdam, 1989, pp. 643–668. DOI:10.1016/B978-0-08-096701-1.00057-4.
- P. K. S. Mural, G. Madras and S. Bose, Polymeric membranes derived from immiscible blends with hierarchical porous structures, tailored bio-interfaces and enhanced flux: Potential and key challenges, Nano-Struct. Nano-Objects, 2018, 14, 149–165, DOI:10.1016/j.nanoso.2018.02.002.
- V. Arrighi, J. M. G. Cowie, S. Fuhrmann and A. Youssef, Miscibility Criterion in Polymer Blends and its Determination, Encyclopedia of Polymer Blends, 2010, 153–198, DOI:10.1002/9783527805204.ch5.
- M. L. Di Lorenzo and M. Frigione, Compatibilization criteria and procedures for binary blends: A review, J. Polym. Eng., 1997, 17(6), 429–460, DOI:10.1515/POLYENG.1997.17.6.429.
- W. Berger, H. W. Kammer and C. Kummerlőwe, Melt rheology and morphology of polymer blends, Macromol. Chem. Phys., 1984, 8(S19841), 101–108, DOI:10.1002/macp.1984.020081984109.
- E. Fekete, E. Földes and B. Pukánszky, Effect of molecular interactions on the miscibility and structure of polymer blends, Eur. Polym. J., 2005, 41(4), 727–736, DOI:10.1016/j.eurpolymj.2004.10.038.
- B. C. Roughton, J. White, K. V. Camarda and R. Gani, Simultaneous Design of Ionic Liquids and Azeotropic Separation Processes, in Computer Aided Chemical Engineering, ed. E. N. Pistikopoulos, M. C. Georgiadis and A. C. Kokossis, Elsevier, 2011, pp. 1578–1582. DOI:10.1016/B978-0-444-54298-4.50094-5.
- J. Fink, Chapter 21 - Dispersions, emulsions, and foams, in Petroleum Engineer's Guide to Oil Field Chemicals and Fluids (Third Edition), ed. J. Fink, Gulf Professional Publishing, 2021, pp. 907–940. DOI:10.1016/B978-0-323-85438-2.00021-9.
- N. Z. Tomić and A. D. Marinković, Chapter 4 - Compatibilization of polymer blends by the addition of graft copolymers, in Compatibilization of Polymer Blends, ed. A. R. Ajitha and S. Thomas, Elsevier, 2020, pp. 103–144. DOI:10.1016/B978-0-12-816006-0.00004-9.
- J. Chen, C. Rong, T. Lin, Y. Chen, J. Wu, J. You, H. Wang and Y. Li, Stable Co-Continuous PLA/PBAT Blends Compatibilized by Interfacial Stereocomplex Crystallites: Toward Full Biodegradable Polymer Blends with Simultaneously Enhanced Mechanical Properties and Crystallization Rates, Macromolecules, 2021, 54(6), 2852–2861, DOI:10.1021/acs.macromol.0c02861.
- S. Tang, R. Zhang, F. Liu and X. Liu, Hansen solubility parameters of polyglycolic acid and interaction parameters between polyglycolic acid and solvents, Eur. Polym. J., 2015, 72, 83–88, DOI:10.1016/j.eurpolymj.2015.09.009.
- Y. Wu, C. Wang, M. Xie and S. Hu, Polybutylene adipate terephthalate/polylactic acid interface enhanced compatibilization and its bead-foaming characteristics, Int. J. Biol. Macromol., 2024, 279, 135221, DOI:10.1016/j.ijbiomac.2024.135221.
- M. Wahbi, Q. Litke, D. Levin, S. Liu, K. J. De France and M. Kontopoulou, Compatibilization of PLA/PBAT blends with epoxidized canola oil for 3D printing applications, Mater. Adv., 2024, 5(12), 5194–5203, 10.1039/D4MA00233D.
- Y. Han, J. Shi, L. Mao, Z. Wang and L. Zhang, Improvement of Compatibility and Mechanical Performances of PLA/PBAT Composites with Epoxidized Soybean Oil as Compatibilizer, Ind. Eng. Chem. Res., 2020, 59(50), 21779–21790, DOI:10.1021/acs.iecr.0c04285.
- P. Nguyen-Tri, P. Ghassemi, P. Carriere, S. Nanda, A. A. Assadi and D. D. Nguyen, Recent Applications of Advanced Atomic Force Microscopy in Polymer Science: A Review, Polymers, 2020, 12(5), 1142, DOI:10.3390/polym12051142.
- D. W. Kim, S. Hwang, S. Hong and E.-C. Lee, Surface Morphology of Polyphenylsilsesquioxanes/Hydroxyl-Functionalized Polystyrene Blends Investigated by Atomic Force Microscopy, Polym. J., 2000, 32, 531–536, DOI:10.1295/POLYMJ.32.531.
- Y. Kikkawa, T. Suzuki, M. Kanesato, Y. Doi and H. Abe, Effect of phase structure on enzymatic degradation in poly(L-lactide)/atactic poly(3-hydroxybutyrate) blends with different miscibility, Biomacromolecules, 2009, 10(4), 1013–1018, DOI:10.1021/bm900117j.
- M. Harada, T. Suzuki, M. Ohya, D. Kawaguchi, A. Takano and Y. Matsushita, Novel miscible polymer blend of poly(4-trimethylsilylstyrene) and polyisoprene, Macromolecules, 2005, 38, 1868–1873, DOI:10.1021/MA048527+.
- M. Pracella, B. Bresci and C. Nicolardi, Phase behaviour and morphology of polymer/liquid crystal blends, Liq. Cryst., 1993, 14, 881–888, DOI:10.1080/02678299308027765.
- M. A. Treece and J. P. Oberhauser, Processing of polypropylene–clay nanocomposites: Single-screw extrusion with in-line supercritical carbon dioxide feed versus twin-screw extrusion, J. Appl. Polym. Sci., 2007, 103(2), 884–892, DOI:10.1002/app.25226.
- J. M. Keen, C. Martin, A. Machado, H. Sandhu, J. W. McGinity and J. C. DiNunzio, Investigation of process temperature and screw speed on properties of a pharmaceutical solid dispersion using corotating and counter-rotating twin-screw extruders, J. Pharm. Pharmacol., 2014, 66(2), 204–217, DOI:10.1111/jphp.12106.
- K. Becker, S. Salar-Behzadi and A. Zimmer, Solvent-Free Melting Techniques for the Preparation of Lipid-Based Solid Oral Formulations, Pharm. Res., 2015, 32(5), 1519–1545, DOI:10.1007/s11095-015-1661-y.
- I. Khan, M. Mansha and M. A. Jafar Mazumder, Polymer Blends, in Functional Polymers, ed. M. A. Jafar Mazumder, H. Sheardown and A. Al-Ahmed, Springer International Publishing, Cham, 2019, pp. 513–549. DOI:10.1007/978-3-319-95987-0_16.
- S. S. Sulistiawan, K. Sadeghi, R. Kumar, D. Kim and J. Seo, In Situ Reactive Extrusion of LDPE Films with Methacrylated Pyrogallol for Antimicrobial and Antioxidant Active Packaging, Polymers, 2025, 17(3), 325, DOI:10.3390/polym17030325.
- D. Nath, E. Pesaranhajiabbas, F. Jahangiri, A. Surendren, A. K. Pal, A. Rodriguez-Uribe, M. Misra and A. K. Mohanty, Maleation of Biodegradable Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Reactive Extrusion: Effect of Initiator Concentration and a Chain Extender on Grafting Percentage and Thermal and Rheological Properties, ACS Omega, 2024, 9(51), 50175–50187, DOI:10.1021/acsomega.4c03477.
- N. Ergun, M. Oksuz and A. Ekinci, Enhancing Mechanical and Thermal Performance of Recycled PA6/PP Blends: Chain Extension and Carbon Fiber Reinforcement Synergy, Materials, 2025, 18(5), 1027, DOI:10.3390/ma18051027.
- S. Agarwal, J. H. Wendorff and A. Greiner, Use of electrospinning technique for biomedical applications, Polymer, 2008, 49(26), 5603–5621, DOI:10.1016/j.polymer.2008.09.014.
- J. Xing, M. Zhang, X. Liu, C. Wang, N. Xu and D. Xing, Multi-material electrospinning: from methods to biomedical applications, Mater. Today Bio, 2023, 21, 100710, DOI:10.1016/j.mtbio.2023.100710.
- H. Ramezani Dana and F. Ebrahimi, Synthesis, properties, and applications of polylactic acid-based polymers, Polym. Eng. Sci., 2023, 63(1), 22–43, DOI:10.1002/pen.26193.
- M. Manfroni, A. Coatti, M. Soccio, V. Siracusa, E. Boanini, E. Salatelli and N. Lotti, Eco-design of biobased poly(butylene succinate-b-pentamethylene 2,5-furanoate) copolymers with optimized mechanical, thermal and barrier properties for flexible food-packaging, Eur. Polym. J., 2025, 225, 113728, DOI:10.1016/j.eurpolymj.2025.113728.
- L. Chel-Guerrero, E. Pérez-Pacheco, J. C. Canto Pinto, V. M. Moo Huchin, I. A. Estrada Mota and R. J. Estrada León, Thermoplastic Starch (TPS)–Cellulosic Fibers Composites: Mechanical Properties and Water Vapor Barrier: A Review, in Composites from Renewable and Sustainable Materials, ed. M. Poletto, IntechOpen, Rijeka, 2016. DOI:10.5772/65397.
- R. Muthuraj, M. Misra and A. K. Mohanty, Biodegradable compatibilized polymer blends for packaging applications: A literature review, J. Appl. Polym. Sci., 2018, 135(24), 45726, DOI:10.1002/app.45726.
- R. R. N. Sailaja and M. Chanda, Use of maleic anhydride–grafted polyethylene as compatibilizer for HDPE–tapioca starch blends: Effects on mechanical properties, J. Appl. Polym. Sci., 2001, 80(6), 863–872, DOI:10.1002/1097-4628(20010509)80:6<863::AID-APP1164>3.0.CO;2-R.
- S. Wang, J. Yu and J. Yu, Compatible thermoplastic starch/polyethylene blends by one-step reactive extrusion, Polym. Int., 2005, 54(2), 279–285, DOI:10.1002/pi.1668.
- P. Ma, A. B. Spoelstra, P. Schmit and P. J. Lemstra, Toughening of poly (lactic acid) by poly (β-hydroxybutyrate-co-β-hydroxyvalerate) with high β-hydroxyvalerate content, Eur. Polym. J., 2013, 49(6), 1523–1531, DOI:10.1016/j.eurpolymj.2013.01.016.
- A. Z. Naser, I. Deiab, F. Defersha and S. Yang, Expanding Poly(lactic acid) (PLA) and Polyhydroxyalkanoates (PHAs) Applications: A Review on Modifications and Effects, Polymers, 2021, 13(23), 4271, DOI:10.3390/polym13234271.
- G. I. C. Righetti, F. Faedi and A. Famulari, Embracing Sustainability: The World of Bio-Based Polymers in a Mini Review, Polymers, 2024, 16(7), 950, DOI:10.3390/polym16070950.
- L. Gao and A. D. Drozdov, Exploring the performance of bio-based PLA/PHB blends: A comprehensive analysis, Polym. Renewable Resour., 2024, 15(3), 358–374, DOI:10.1177/20412479241266954.
- M. Kervran, C. Vagner, M. Cochez, M. Ponçot, M. R. Saeb and H. Vahabi, Thermal degradation of polylactic acid (PLA)/polyhydroxybutyrate (PHB) blends: A systematic review, Polym. Degrad. Stab., 2022, 201, 109995, DOI:10.1016/j.polymdegradstab.2022.109995.
- M. Qahtani, F. Wu, M. Misra, S. Gregori, D. F. Mielewski and A. K. Mohanty, Experimental Design of Sustainable 3D-Printed Poly(Lactic Acid)/Biobased Poly(Butylene Succinate) Blends via Fused Deposition Modeling, ACS Sustainable Chem. Eng., 2019, 7(17), 14460–14470, DOI:10.1021/acssuschemeng.9b01830.
- I. Armentano, E. Fortunati, N. Burgos, F. Dominici, F. Luzi, S. Fiori, A. Jiménez, K. Yoon, J. Ahn, S. Kang and J. M. Kenny, Processing and characterization of plasticized PLA/PHB blends for biodegradable multiphase systems, eXPRESS Polym. Lett., 2015, 9, 583–596 CrossRef CAS.
- X. Wang, S. Peng, H. Chen, X. Yu and X. Zhao, Mechanical properties, rheological behaviors, and phase morphologies of high-toughness PLA/PBAT blends by in situ reactive compatibilization, Composites, Part B, 2019, 173, 107028, DOI:10.1016/j.compositesb.2019.107028.
- M. Aldas, J. M. Ferri, D. L. Motoc, L. Peponi, M. P. Arrieta and J. López-Martínez, Gum Rosin as a Size Control Agent of Poly(Butylene Adipate-Co-Terephthalate) (PBAT) Domains to Increase the Toughness of Packaging Formulations Based on Polylactic Acid (PLA), Polymers, 2021, 13(12), 1913, DOI:10.3390/polym13121913.
- B. M. Trinh, D. T. Tadele and T. H. Mekonnen, Robust and high barrier thermoplastic starch – PLA blend films using starch-graft-poly(lactic acid) as a compatibilizer, Mater. Adv., 2022, 3(15), 6208–6221, 10.1039/D2MA00501H.
- J. Resch, M. Kreutzbruck and C. Bonten, Modification of PHBV-blends with a biodegradable plasticizer, AIP Conf. Proc., 2020, 2289(1), 020059, DOI:10.1063/5.0028771.
- K. G. Patel, R. K. Maynard, L. S. I. V. Ferguson, M. L. Broich, II, J. C. Bledsoe, C. C. Wood, G. H. Crane, J. A. Bramhall, J. M. Rust, A. Williams-Rhaesa and J. J. Locklin, Experimentally Determined Hansen Solubility Parameters of Biobased and Biodegradable Polyesters, ACS Sustainable Chem. Eng., 2024, 12(6), 2386–2393, DOI:10.1021/acssuschemeng.3c07284.
- S. Su, Prediction of the Miscibility of PBAT/PLA Blends, Polymers, 2021, 13, 2339, DOI:10.3390/polym13142339.
- Y. Ding, W. Feng, D. Huang, B. Lu, P. Wang, G. Wang and J. Ji, Compatibilization of immiscible PLA-based biodegradable polymer blends using amphiphilic di-block copolymers, Eur. Polym. J., 2019, 118, 45–52, DOI:10.1016/j.eurpolymj.2019.05.036.
- R. C. C. Domingues, C. C. Pereira and C. P. Borges, Morphological control and properties of poly(lactic acid) hollow fibers for biomedical applications, J. Appl. Polym. Sci., 2017, 134(47), 45494, DOI:10.1002/app.45494.
- J. S. Choi and W. H. Park, Effect of biodegradable plasticizers on thermal and mechanical properties of poly(3-hydroxybutyrate), Polym. Test., 2004, 23(4), 455–460, DOI:10.1016/j.polymertesting.2003.09.005.
- K. W. Meereboer, A. K. Pal, M. Misra and A. K. Mohanty, Sustainable PHBV/Cellulose Acetate Blends: Effect of a Chain Extender and a Plasticizer, ACS Omega, 2020, 5(24), 14221–14231, DOI:10.1021/acsomega.9b03369.
- M. Imaizumi, T. Nagata, Y. Goto, Y. Okino, T. Takahashi and K. Koyama, Solubility Parameters of Biodegradable Polymers from Turbidimetric Titrations, Kobunshi Ronbunshu, 2005, 62(9), 438–440, DOI:10.1295/koron.62.438.
- D. Li, B. Zhang, S. Wang, C. E. Zhou, C. Sun and Q. Zha, A systematic study of dyeing polybutylene succinate fibres, Color. Technol., 2020, 136(1), 87–96, DOI:10.1111/cote.12448.
- A. Cascales, C. Pavon, S. Ferrandiz and J. López-Martínez, Evaluation of Thermoplastic Starch Contamination in the Mechanical Recycling of High-Density Polyethylene, Recycling, 2024, 9(3), 33, DOI:10.3390/recycling9030033.
- O. I. Olejnik, A. Masek and J. J. M. Zawadziłło, Processability and Mechanical Properties of Thermoplastic Polylactide/Polyhydroxybutyrate (PLA/PHB) Bioblends, Materials, 2021, 14(4), 898 CrossRef CAS PubMed.
- R. Plavec, S. Hlaváčiková, L. Omaníková, J. Feranc, Z. Vanovčanová, K. Tomanová, J. Bočkaj, J. Kruželák, E. Medlenová, I. Gálisová, L. Danišová, R. Přikryl, S. Figalla, V. Melčová and P. Alexy, Recycling possibilities of bioplastics based on PLA/PHB blends, Polym. Test., 2020, 92, 106880, DOI:10.1016/j.polymertesting.2020.106880.
- I. Ohkoshi, H. Abe and Y. Doi, Miscibility and solid-state structures for blends of poly[(S)-lactide] with atactic poly[(R,S)-3-hydroxybutyrate], Polymer, 2000, 41(15), 5985–5992, DOI:10.1016/S0032-3861(99)00781-8.
- M. S. Popa, A. N. Frone and D. M. Panaitescu, Polyhydroxybutyrate blends: A solution for biodegradable packaging?, Int. J. Biol. Macromol., 2022, 207, 263–277, DOI:10.1016/j.ijbiomac.2022.02.185.
- M. L. Iglesias-Montes, M. Soccio, V. Siracusa, M. Gazzano, N. Lotti, V. P. Cyras and L. B. Manfredi, Chitin
Nanocomposite Based on Plasticized Poly(lactic acid)/Poly(3-hydroxybutyrate) (PLA/PHB) Blends as Fully Biodegradable Packaging Materials, Polymers, 2022, 14(15), 3177, DOI:10.3390/polym14153177.
- A. D’Anna, R. Arrigo and A. Frache, PLA/PHB Blends: Biocompatibilizer Effects, Polymers, 2019, 11(9), 1416, DOI:10.3390/polym11091416.
- D. P. Ura and U. Stachewicz, The Significance of Electrical Polarity in Electrospinning: A Nanoscale Approach for the Enhancement of the Polymer Fibers’ Properties, Macromol. Mater. Eng., 2022, 307(5), 2100843, DOI:10.1002/mame.202100843.
- J. S. C. Lo, X. Chen, S. Chen, Y. Miao, W. A. Daoud, C. Y. Tso, I. Firdous, B. J. Deka and C. S. K. Lin, Fabrication of biodegradable PLA-PHBV medical textiles via electrospinning for healthcare apparel and personal protective equipment, Sustainable Chem. Pharm., 2024, 39, 101536, DOI:10.1016/j.scp.2024.101536.
- C. Brütting, J. Dreier, C. Bonten and H. Ruckdäschel, Biobased Immiscible Polylactic Acid (PLA): Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) Blends: Impact of Rheological and Non-isothermal Crystallization on the Bead Foaming Behavior, J. Polym. Environ., 2024, 32(9), 4182–4195, DOI:10.1007/s10924-024-03186-9.
- S. Gasmi, M. K. Hassan and A. S. Luyt, Crystallization and dielectric behaviour of PLA and PHBV in PLA/PHBV blends and PLA/PHBV/TiO2 nanocomposites, eXPRESS Polym. Lett., 2019, 13(2), 199–212, DOI:10.3144/expresspolymlett.2019.16.
- C. Brütting, J. Dreier, C. Bonten, V. Altstädt and H. Ruckdäschel, Sustainable Immiscible Polylactic Acid (PLA) and Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) Blends: Crystallization and Foaming Behavior, ACS Sustainable Chem. Eng., 2023, 11(17), 6676–6687, DOI:10.1021/acssuschemeng.3c00199.
- J. Novak, L. Behalek, J. Hlozek, M. Boruvka, P. Brdlik and P. Lenfeld, Non-isothermal crystallization kinetics of biocomposites based on PLA/PHBV and spent coffee grounds, J. Therm. Anal. Calorim., 2024, 150, 1225–1243, DOI:10.1007/s10973-024-13480-2.
- J. S. C. Lo, X. Chen, S. Chen, Y. Miao, W. A. Daoud, C. Y. Tso, I. Firdous, B. J. Deka and C. S. K. Lin, Fabrication of biodegradable PLA-PHBV medical textiles via electrospinning for healthcare apparel and personal protective equipment, Sustainable Chem. Pharm., 2024, 39, 101536, DOI:10.1016/j.scp.2024.101536.
- H. Pouriman, R. Lin, K. Graham and K. Jayaraman, Evaluating the grafting of Maleic anhydride with Poly(lactic acid) (PLA) and Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) via batch compounding, Mater. Today: Proc., 2023 DOI:10.1016/j.matpr.2023.03.526.
- S. M. K. Hasan, M. K. Hossain, S. N. Shaily, H. J. Harrigan and T. Mickens, Mechanical, Thermal, and Morphological Study of Nanoclay Reinforced Bio-based Poly Lactic Acid/Poly (3-hydroxy butyrate co-3-valerate) (PLA/PHBV) Blend, Global J. Eng. Sci., 2020, 6(2) DOI:10.33552/GJES.2020.06.000631.
- M. A. Vigil Fuentes, S. Thakur, F. Wu, M. Misra, S. Gregori and A. K. Mohanty, Study on the 3D printability of poly(3-hydroxybutyrate-co-3-hydroxyvalerate)/poly(lactic acid) blends with chain extender using fused filament fabrication, Sci. Rep., 2020, 10(1), 11804, DOI:10.1038/s41598-020-68331-5.
- E. Hernández-García, M. Vargas and A. Chiralt, Effect of active phenolic acids on properties of PLA-PHBV blend films, Food Packag. Shelf Life, 2022, 33, 100894, DOI:10.1016/j.fpsl.2022.100894.
- M. P. Arrieta, M. Perdiguero, S. Fiori, J. M. Kenny and L. Peponi, Biodegradable electrospun PLA-PHB fibers plasticized with oligomeric lactic acid, Polym. Degrad. Stab., 2020, 179, 109226, DOI:10.1016/j.polymdegradstab.2020.109226.
- J. Y. Boey, U. Kong, C. K. Lee, G. K. Lim, C. W. Oo, C. K. Tan, C. Y. Ng, A. A. Azniwati and G. S. Tay, The effect of filler loading, biological treatment, and bioplastic blend ratio on flexural and impact properties of blended bioplastic reinforced with spent coffee ground, Polym. Eng. Sci., 2024, 64(7), 3319–3333, DOI:10.1002/pen.26772.
- M. Musioł, J. Rydz, H. Janeczek, J. Andrzejewski, M. Cristea, K. Musioł, M. Kampik and M. Kowalczuk, (Bio)degradable Biochar Composites of PLA/P(3HB-co-4HB) Commercial Blend for Sustainable Future—Study on Degradation and Electrostatic Properties, Polymers, 2024, 16, 2331, DOI:10.3390/polym16162331.
- S. Jurczyk, J. Andrzejewski, A. Piasecki, M. Musioł, J. Rydz and M. Kowalczuk, Mechanical and Rheological Evaluation of Polyester-Based Composites Containing Biochar, Polymers, 2024, 16(9), 1231, DOI:10.3390/polym16091231.
- M. Dammak, Y. Fourati, Q. Tarrés, M. Delgado-Aguilar, P. Mutjé and S. J. I. C. Boufi, and Products, Blends of PBAT with plasticized starch for packaging applications: Mechanical properties, rheological behaviour and biodegradability, Ind. Crops Prod., 2020, 144, 112061, DOI:10.1016/j.indcrop.2019.112061.
- M. Meng, S. Wang, M. Xiao and Y. Meng, Recent Progress in Modification and Preparations of the Promising Biodegradable Plastics: Polylactide and Poly(butylene adipate-co-terephthalate), Sustainable Polym. Energy, 2023, 1(1), 1–43, DOI:10.35534/spe.2023.10006.
- W. Liu, Y. Wang, S. Xiang and H. Liu, Unveiling the effect of enhanced interfacial compatibility on the mechanical properties of PLA/PBAT blends, Polymer, 2024, 296, 126815, DOI:10.1016/j.polymer.2024.126815.
- A. Raj, M. Yousfi, K. Prashantha and C. Samuel, Morphologies, Compatibilization and Properties of Immiscible PLA-Based Blends with Engineering Polymers: An Overview of Recent Works, Polymers, 2024, 16(13), 1776, DOI:10.3390/polym16131776.
- X. Ai, X. Li, Y. Yu, H. Pan, J. Yang, D. Wang, H. Yang, H. Zhang and L. Dong, The Mechanical, Thermal, Rheological and Morphological Properties of PLA/PBAT Blown Films by Using Bis(tert-butyl dioxy isopropyl) Benzene as Crosslinking Agent, Polym. Eng. Sci., 2019, 59(S1), E227–E236, DOI:10.1002/pen.24927.
- M. Sun, L. Zhang and C. Li, Modified cellulose nanocrystals based on SI-ATRP for enhancing interfacial compatibility and mechanical performance of biodegradable PLA/PBAT blend, Polym. Compos., 2022, 43(6), 3753–3764, DOI:10.1002/pc.26653.
- J. Zhang, P. Li, Y. Li, M. Luo, Z. Yan, T. Wang, Q. Fu, X. Gao and J. Zhang, Preparation of PLA/PBAT blends with high performance via the synergistic effect of high mold temperature and strong shear field, Polymer, 2024, 296, 126795, DOI:10.1016/j.polymer.2024.126795.
- A. Folino, D. Pangallo and P. S. Calabrò, Assessing bioplastics biodegradability by standard and research methods: Current trends and open issues, J. Environ. Chem. Eng., 2023, 11(2), 109424, DOI:10.1016/j.jece.2023.109424.
- Y. Qiu, P. Wang, L. Zhang, C. Li, J. Lu and L. Ren, Enhancing biodegradation efficiency of PLA/PBAT-ST20 bioplastic using thermophilic bacteria co-culture system: New insight from structural characterization, enzyme activity, and metabolic pathways, J. Hazard. Mater., 2024, 477, 135426, DOI:10.1016/j.jhazmat.2024.135426.
- P. D. Dissanayake, P. A. Withana, M. K. Sang, Y. Cho, J. Park, D. X. Oh, S. X. Chang, C. S. K. Lin, M. S. Bank, S. Y. Hwang and Y. S. Ok, Effects of biodegradable poly(butylene adipate-co-terephthalate) and poly(lactic acid) plastic degradation on soil ecosystems, Soil Use Manage., 2024, 40(2), e13055, DOI:10.1111/sum.13055.
- W. Peng, R. Nie, F. Lü, H. Zhang and P. He, Biodegradability of PBAT/PLA coated paper and bioplastic bags under anaerobic digestion, Waste Manage., 2024, 174, 218–228, DOI:10.1016/j.wasman.2023.11.037.
- Q. Wang, X. Zou, S. Kang, Y. Wang and Z. Li, Degradation of polylactic acid/polybutylene adipate-co-terephthalate blend by Papiliotrema laurentii S2P4P isolated from agricultural soils, Polym. Degrad. Stab., 2024, 227, 110855, DOI:10.1016/j.polymdegradstab.2024.110855.
- J. Benninga, G. M. R. Lima, G. Érsek, G. Portale, R. Folkersma, V. S. D. Voet and K. Loos, Enzymatic Degradation of Poly(Butylene Adipate-co-Terephthalate)/Poly(Lactic Acid) Blends, J. Polym. Sci., 2024 DOI:10.1002/pol.20240785.
- G. F. Toledo, G. F. Schutz, L. Marangoni Júnior and R. P. Vieira, Thermo-pressed blend films of poly(lactic acid)/poly(butylene adipate-co-terephthalate) with polylimonene for sustainable active food packaging, Sustainable Mater. Technol., 2024, 41, e01099, DOI:10.1016/j.susmat.2024.e01099.
- W. Yu, M. Li, W. Lei and Y. Chen, FDM 3D Printing and Properties of PBAT/PLA Blends, Polymers, 2024, 16(8), 1140, DOI:10.3390/polym16081140.
- L. L. R. L. de Castro, L. G. L. Silva, I. R. Abreu, C. J. F. Braz, S. C. S. Rodrigues, R. S. d. R. Moreira-Araújo, R. Folkersma, L. H. de Carvalho, R. Barbosa and T. S. Alves, Biodegradable PBAT/PLA blend films incorporated with turmeric and cinnamomum powder: A potential alternative for active food packaging, Food Chem., 2024, 439, 138146, DOI:10.1016/j.foodchem.2023.138146.
- C.-H. Tsou, H. Luo, S.-M. Lin, C. Preuksarattanawut, P. Potiyaraj, C.-S. Wu, F.-F. Ge, J. Du and X. Wei, Eco-friendly enhancement of poly(lactic acid)/poly(butylene adipate-co-terephthalate) bridgeable composites using natural cotton stalk: A novel approach to improved mechanical, barrier properties, and compatibility, Polym. Eng. Sci., 2024, 64(10), 4786–4800, DOI:10.1002/pen.26881.
- G. S. Sudha, N. R. Aswathy, C. B. Mahesh and K. M. Aswini, Investigating the reinforcing effect of jute fiber in a PLA & PBAT biopolymer blend matrix for advanced engineering applications: Enhancing sustainability with bioresources, Mater. Today Commun., 2024, 40, 109960, DOI:10.1016/j.mtcomm.2024.109960.
- F. C. Nunes, K. C. Ribeiro, F. A. Martini, B. R. Barrioni, J. P. F. Santos and B. Melo Carvalho, PBAT/PLA/cellulose nanocrystals biocomposites compatibilized with polyethylene grafted maleic anhydride (PE-g-MA), J. Appl. Polym. Sci., 2021, 138(45), 51342, DOI:10.1002/app.51342.
- D. Jubinville, M. Awad, H.-S. Lee and T. H. Mekonnen, Effect of Compatibilizers on the Physico-mechanical Properties of a Poly(Lactic Acid)/Poly(Butylene Adipate-co-terephthalate) Matrix with Rice Straw Micro-particle Fillers, J. Polym. Environ., 2024, 32, 5857–5872, DOI:10.1007/s10924-024-03314-5.
- S. Chuayjuljit, T. Leejarkpai and P. Chaiwutthinan, Effect of accelerated weathering on the properties of poly(lactic acid)/poly(butylene adipate-co-terephthalate) blend and composite containing wood flour and wollastonite, Prog. Rubber, Plast. Recycl. Technol., 2023, 40(3), 268–286, DOI:10.1177/14777606231219641.
- P. Zytner, A. K. Pal, A. K. Mohanty and M. Misra, Performance evaluation of biodegradable polymer PHBV and PBAT blends with adjustable melt flow behaviour, heat deflection temperature, and morphological transition, Can. J. Chem. Eng., 2024, 102(8), 2805–2817, DOI:10.1002/cjce.25235.
- F. Jahangiri, A. K. Mohanty, A. K. Pal, S. Shankar, A. Rodriguez-Uribe, R. Clemmer, S. Gregori and M. Misra, PHBV coating on biodegradable plastic sheet: Effect of coating on morphological, mechanical and barrier properties, Prog. Org. Coat., 2024, 189, 108270, DOI:10.1016/j.porgcoat.2024.108270.
- A. S. Narmon, A. Dewaele, K. Bruyninckx, B. F. Sels, P. Van Puyvelde and M. Dusselier, Boosting PLA melt strength by controlling the chirality of co-monomer incorporation, Chem. Sci., 2021, 12(15), 5672–5681, 10.1039/d1sc00040c.
- A. K. Pal, F. Wu, M. Misra and A. K. Mohanty, Reactive extrusion of sustainable PHBV/PBAT-based nanocomposite films with organically modified nanoclay for packaging applications: Compression moulding vs. cast film extrusion, Composites, Part B, 2020, 198, 108141, DOI:10.1016/j.compositesb.2020.108141.
- M. Fernandes, A. F. Salvador and A. A. Vicente, Biodegradation of PHB/PBAT films and isolation of novel PBAT biodegraders from soil microbiomes, Chemosphere, 2024, 362, 142696, DOI:10.1016/j.chemosphere.2024.142696.
- A. R. M. Costa, E. N. Ito, L. H. Cavalho and E. L. Canedo, Non-isothermal melt crystallization kinetics of poly(3-hydroxybutyrate), poly(butylene adipate-co-terephthalate) and its mixture, Polímeros, 2019, 29(1), e2019006, DOI:10.1590/0104-1428.11217.
- V. C. Beber, S. De Barros, M. D. Banea, M. Brede, L. H. De Carvalho, R. Hoffmann, A. R. M. Costa, E. B. Bezerra, I. D. S. Silva, K. Haag, K. Koschek and R. M. R. Wellen, Effect of Babassu Natural Filler on PBAT/PHB Biodegradable Blends: An Investigation of Thermal, Mechanical, and Morphological Behavior, Materials, 2018, 11(5), 820, DOI:10.3390/ma11050820.
- D. Hlotse, W. Mhike, V. Ojijo, M. B. Shongwe and M. J. John, Compostable Materials From PHA Based Blends and Composites, Reference Module in Materials Science and Materials Engineering, 2024 DOI:10.1016/B978-0-323-95486-0.00053-3 , Elsevier.
- S.-J. Xiong, B. Pang, S.-J. Zhou, M.-K. Li, S. Yang, Y.-Y. Wang, Q. Shi, S.-F. Wang, T.-Q. Yuan and R.-C. Sun, Economically Competitive Biodegradable PBAT/Lignin Composites: Effect of Lignin Methylation and Compatibilizer, ACS Sustainable Chem. Eng., 2020, 8(13), 5338–5346, DOI:10.1021/acssuschemeng.0c00789.
- V. C. Beber, S. de Barros, M. D. Banea, M. Brede, L. H. de Carvalho, R. Hoffmann, A. R. M. Costa, E. B. Bezerra, I. D. S. Silva, K. Haag, K. Koschek and R. M. R. Wellen, Effect of Babassu Natural Filler on PBAT/PHB Biodegradable Blends: An Investigation of Thermal, Mechanical, and Morphological Behavior, Materials, 2018, 11(5), 820, DOI:10.3390/ma11050820.
- A. Gupta, L. Lolic and T. H. Mekonnen, Reactive extrusion of highly filled, compatibilized, and sustainable PHBV/PBAT – Hemp residue biocomposite, Composites, Part A, 2022, 156, 106885, DOI:10.1016/j.compositesa.2022.106885.
- A. K. Pal, M. Misra and A. K. Mohanty, Silane treated starch dispersed PBAT/PHBV-based composites: Improved barrier performance for single-use plastic alternatives, Int. J. Biol. Macromol., 2023, 229, 1009–1022, DOI:10.1016/j.ijbiomac.2022.12.141.
- Y. Dasan, A. Bhat and F. J. C. p. Ahmad, Polymer blend of PLA/PHBV based bionanocomposites reinforced with nanocrystalline cellulose for potential application as packaging material, Carbohydr. Polym., 2017, 157, 1323–1332, DOI:10.1016/j.carbpol.2016.11.012.
- L. Botta, V. Titone, M. C. Mistretta, F. P. La Mantia, A. Modica, M. Bruno, F. Sottile and F. Lopresti, PBAT Based Composites Reinforced with Microcrystalline Cellulose Obtained from Softwood Almond Shells, Polymers, 2021, 13(16), 2643, DOI:10.3390/polym13162643.
- O. Yolacan and S. Deniz, Effects of different biopolymers and additives on mechanical and barrier properties of Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) blend films, J. Dispersion Sci. Technol., 2024, 46(10), 1530–1539, DOI:10.1080/01932691.2024.2334028.
- B. Melendez-Rodriguez, C. Prieto, M. Pardo-Figuerez, I. Angulo, A. I. Bourbon, I. R. Amado, M. A. Cerqueira, L. M. Pastrana, L. H. Hilliou, A. A. Vicente, L. Cabedo and J. M. Lagaron, Multilayer Film Comprising Polybutylene Adipate Terephthalate and Cellulose Nanocrystals with High Barrier and Compostable Properties, Polymers, 2024, 16(15), 2095, DOI:10.3390/polym16152095.
- K. Martinez Villadiego, M. J. Arias Tapia, J. Useche and D. Escobar Macías, Thermoplastic Starch (TPS)/Polylactic Acid (PLA) Blending Methodologies: A Review, J. Polym. Environ., 2022, 30(1), 75–91, DOI:10.1007/s10924-021-02207-1.
- A. Yamaguchi, S. Arai and N. Arai, Design strategy for blends of biodegradable polyester and thermoplastic starch based on a molecular dynamics study of the phase-separated interface, Carbohydr. Polym., 2024, 333, 122005, DOI:10.1016/j.carbpol.2024.122005.
- W. Qin, Y. Qiu, H. He, B. Guo and P. Li, Pyrogallic acid–compatibilized polylactic acid/thermoplastic starch blend produced via one-step twin-screw extrusion, Int. J. Biol. Macromol., 2024, 276, 133758, DOI:10.1016/j.ijbiomac.2024.133758.
- V. Sessini, V. Salaris, V. Oliver-Cuenca, A. Tercjak, S. Fiori, D. López, J. M. Kenny and L. Peponi, Thermally-Activated Shape Memory Behavior of Biodegradable Blends Based on Plasticized PLA and Thermoplastic Starch, Polymers, 2024, 16(8), 1107, DOI:10.3390/polym16081107.
- L. Songtipya, E. Kalkornsurapranee, P. Songtipya, T. Sengsuk, R. Promsung, A. Chuaybamrung and J. Johns, Enhancing Thermo-mechanical Properties of Thermoplastic Starch/Natural Rubber Blends Through the Synergistic Combination of PEG and Modified Natural Rubber, J. Polym. Environ., 2024, 32(4), 1868–1878, DOI:10.1007/s10924-023-03086-4.
- C. A. Ávila-Orta, C. A. Covarrubias-Gordillo, H. A. Fonseca-Florido, L. Melo-López, R. Radillo-Ruíz and E. Gutiérrez-Montiel, PLA/modified-starch blends and their application for the fabrication of non-woven fabrics by melt-blowing, Carbohydr. Polym., 2023, 316, 120975, DOI:10.1016/j.carbpol.2023.120975.
- A. Fonseca-García, B. H. Osorio, R. Y. Aguirre-Loredo, H. L. Calambas and C. Caicedo, Miscibility study of thermoplastic starch/polylactic acid blends: Thermal and superficial properties, Carbohydr. Polym., 2022, 293, 119744, DOI:10.1016/j.carbpol.2022.119744.
- S. Chauhan, N. Raghu and A. Raj, Effect of maleic anhydride grafted polylactic acid concentration on mechanical and thermal properties of thermoplasticized starch filled polylactic acid blends, Polym. Polym. Compos., 2021, 29(9_suppl), S400–S410, DOI:10.1177/09673911211004194.
- N. Perumal, S. Sreekantan, Z. A. A. Hamid, A. Rusli, K. Bhubalan and J. N. Appaturi, Effect of Plasticizer and Compatibilizer on Properties of Polybutylene Adipate-Co-Terephthalate (PBAT) with Acetylated Starch, J. Polym. Environ., 2024, 32(1), 289–302, DOI:10.1007/s10924-023-02964-1.
- H. Eslami, M. Grady and T. H. Mekonnen, Biobased and compostable trilayer thermoplastic films based on poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) and thermoplastic starch (TPS), Int. J. Biol. Macromol., 2022, 220, 385–394, DOI:10.1016/j.ijbiomac.2022.08.079.
- P. Kampeerapappun, N. O-Charoen, P. Dhamvithee and E. Jansri, Biocomposite Based on Polylactic Acid and Rice Straw for Food Packaging Products, Polymers, 2024, 16(8), 1038, DOI:10.3390/polym16081038.
- P. A. V. Freitas, C. González-Martínez and A. Chiralt, Stability and Composting Behaviour of PLA–Starch Laminates Containing Active Extracts and Cellulose Fibres from Rice Straw, Polymers, 2024, 16(11), 1474, DOI:10.3390/polym16111474.
- X. Zhai, J. Han, L. Chang, F. Zhao, R. Zhang, W. Wang and H. Hou, Effects of starch filling on physicochemical properties, functional activities, and release characteristics of PBAT-based biodegradable active films loaded with tea polyphenols, Int. J. Biol. Macromol., 2024, 277, 134505, DOI:10.1016/j.ijbiomac.2024.134505.
- M. Aouay, A. Magnin, C. Lancelon-Pin, J.-L. Putaux and S. Boufi, Mitigating the Water Sensitivity of PBAT/TPS Blends through the Incorporation of Lignin-Containing Cellulose Nanofibrils for Application in Biodegradable Films, ACS Sustainable Chem., 2024, 12(29), 10805–10819, DOI:10.1021/acssuschemeng.4c02245.
- L. F. Silva, P. H. P. M. d. Silveira, A. C. B. Rodrigues, S. N. Monteiro, S. F. Santos, J. P. S. Morais and D. C. Bastos, Cotton incorporated Poly(lactic acid)/thermoplastic Starch Based Composites Used as Flexible Packing for Short Shelf Life Products, Mater. Res., 2024, 27, e20230366, DOI:10.1590/1980-5373-MR-2023-0366.
- X. Tang, J. Tan, Y. Hu, C. Su, Z. Liu, C. Wei, S. Dong and F. Meng, Study on mechanical and thermal properties of coir fibers reinforced thermoplastic starch/poly(butylene adipate-co-terephthalate) composites, J. Appl. Polym. Sci., 2024, e56206, DOI:10.1002/app.56206.
- D. C. McNeill, A. K. Pal, A. K. Mohanty and M. Misra, Injection molding of biodegradable polyester blends filled with mineral and sustainable fillers: Performance evaluation, J. Appl. Polym. Sci., 2024, 141(13), e55166, DOI:10.1002/app.55166.
- M. Barletta, C. Aversa, M. Ayyoob, A. Gisario, K. Hamad, M. Mehrpouya and H. Vahabi, Poly(butylene succinate) (PBS): Materials, processing, and industrial applications, Prog. Polym. Sci., 2022, 132, 101579, DOI:10.1016/j.progpolymsci.2022.101579.
- M. d. R. Salazar-Sánchez, L. I. Delgado-Calvache, J. C. Casas-Zapata, H. S. Villada Castillo and J. F. Solanilla-Duque, Soil Biodegradation of a Blend of Cassava Starch and Polylactic Acid, Ing. Invest., 2022, 42(3), e93710, DOI:10.15446/ing.investig.93710.
- A. Paul, K. Sreedevi, S. S. Sharma and V. N. Anjana, Polylactic Acid (PLA), in Handbook of Biopolymers, ed. S. Thomas, et al., 2023, Springer Nature Singapore, Singapore. pp. 1195–1227. DOI:10.1007/978-981-19-0710-4_44.
- S. Adrar and A. Ajji, Effect of blending sequence and epoxy functionalized compatibilizer on barrier and mechanical properties of PBS and PBS/PLA nanocomposite blown films, J. Vinyl Addit. Technol., 2024, 30(4), 883–894, DOI:10.1002/vnl.22092.
- L. Meng, M. Chen, X. Sun, Z. Li, N. Liu and X. Wang, Tailoring the Microstructure of Biodegradable PLA/PBS Melt-Blown Nonwovens with Enhanced Mechanical Performance by In Situ PBS Fibrils Formation, Ind. Eng. Chem. Res., 2024, 63(29), 13016–13024, DOI:10.1021/acs.iecr.4c00591.
- S. Chuakhao, J. T. Rodríguez, S. Lapnonkawow, G. Kannan, A. J. Müller and S. Suttiruengwong, Formulating PBS/PLA/PBAT blends for biodegradable, compostable packaging: The crucial roles of PBS content and reactive extrusion, Polym. Test., 2024, 132, 108383, DOI:10.1016/j.polymertesting.2024.108383.
- A. Apicella, K. V. Malafeev, P. Scarfato and L. Incarnato, Generation of Microplastics from Biodegradable Packaging Films Based on PLA, PBS and Their Blend in Freshwater and Seawater, Polymers, 2024, 16(16), 2268, DOI:10.3390/polym16162268.
- I. S. Choi, Y. K. Kim, S. H. Hong, H.-J. Seo, S.-H. Hwang, J. Kim and S. K. Lim, Effects of Polybutylene Succinate Content on the Rheological Properties of Polylactic Acid/Polybutylene Succinate Blends and the Characteristics of Their Fibers, Materials, 2024, 17(3), 662, DOI:10.3390/ma17030662.
- H. Peshne, K. P. Das, D. Sharma and B. K. Satapathy, Physico-mechanical Evaluation of Electrospun Nanofibrous Mats of Poly(3-hydroxybutyrate)/Poly(butylene succinate) Blends with Enhanced Swelling-Dynamics and Hydrolytic Degradation-Kinetics Stability for Pliable Scaffold Substrates, J. Polym. Environ., 2024, 32(8), 4046–4067, DOI:10.1007/s10924-023-03174-5.
- J. Kim, H. Yun, S. Won, D. Lee, S. Baek, G. Heo, S. Park, H.-J. Jin and H. W. Kwak, Comparative degradation behavior of polybutylene succinate (PBS), used PBS, and PBS/Polyhydroxyalkanoates (PHA) blend fibers in compost and marine–sediment interfaces, Sustainable Mater. Technol., 2024, 41, e01065, DOI:10.1016/j.susmat.2024.e01065.
- K. Samaniego-Aguilar, E. Sanchez-Safont, I. Pisa-Ripoll, S. Torres-Giner, Y. Flores, J. M. Lagaron, L. Cabedo and J. Gamez-Perez, Performance Enhancement of Biopolyester Blends by Reactive Compatibilization with Maleic Anhydride-Grafted Poly(butylene succinate-co-adipate), Polymers, 2024, 16(16), 2325, DOI:10.3390/polym16162325.
- P. Feijoo, A.
K. Mohanty, A. Rodriguez-Uribe, J. Gámez-Pérez, L. Cabedo and M. Misra, Biodegradable blends from bacterial biopolyester PHBV and bio-based PBSA: Study of the effect of chain extender on the thermal, mechanical and morphological properties, Int. J. Biol. Macromol., 2023, 225, 1291–1305, DOI:10.1016/j.ijbiomac.2022.11.188.
- Z. Ma, T. Yin, Z. Jiang, Y. Weng and C. Zhang, Bio-based epoxidized soybean oil branched cardanol ethers as compatibilizers of polybutylene succinate (PBS)/polyglycolic acid (PGA) blends, Int. J. Biol. Macromol., 2024, 259, 129319, DOI:10.1016/j.ijbiomac.2024.129319.
- N. Rajendran and J. Han, Techno-economic analysis and life cycle assessment of poly (butylene succinate) production using food waste, Waste Manage., 2023, 156, 168–176, DOI:10.1016/j.wasman.2022.11.037.
- Y. Li, Z. Li, S. Sheng, Y. Li, J.-R. Zhong, J. Tan and Y.-F. Zhang, Preparation and properties of rapidly plasticized poly (butylene succinate)/mechanically activated cassava starch biocomposite, Polym. Bull., 2024, 81(7), 6495–6511, DOI:10.1007/s00289-023-05018-7.
- N. Thajai, P. Rachtanapun, S. Thanakkasaranee, W. Punyodom, P. Worajittiphon, Y. Phimolsiripol, N. Leksawasdi, S. Ross, P. Jantrawut and K. Jantanasakulwong, Reactive Blending of Modified Thermoplastic Starch Chlorhexidine Gluconate and Poly(butylene succinate) Blending with Epoxy Compatibilizer, Polymers, 2023, 15(16), 3487, DOI:10.3390/polym15163487.
- P. Pei, Y. Sun, R. Zou, X. Wang, J. Liu, L. Liu, X. Deng, X. Li, M. Yu and S. Li, Comparing four kinds of lignocellulosic biomass for the performance of fiber/PHB/PBS bio-composites, BioResources, 2023, 18(4), 7154–7171, DOI:10.15376/biores.18.4.7154-7171.
- A. Apicella, G. Molinari, V. Gigante, A. Pietrosanto, L. Incarnato, L. Aliotta and A. Lazzeri, Poly(lactic acid) (PLA)/poly(butylene succinate adipate) (PBSA) films with Micro fibrillated cellulose (MFC) and cardanol for packaging applications, Cellulose, 2024, 31, 9173–9190, DOI:10.1007/s10570-024-06127-w.
- J. O. Akindoyo, K. Pickering, M. D. Beg and M. Mucalo, Reactive compatibilization of harakeke fiber-reinforced poly(lactic) acid/polybutylene succinate blend, J. Appl. Polym. Sci., 2024, 141(40), e56030, DOI:10.1002/app.56030.
- N. Ketata, M. Ejday, Y. Grohens, B. Seantier and N. Guermazi, Investigation of the hybridization effect on mechanical properties of natural fiber reinforced biosourced composites, J. Compos. Mater., 2024, 58(17), 1965–1985, DOI:10.1177/00219983241255751.
- M. A. Masanabo, A. Tribot, E. Luoma, J. Virkajärvi, N. Sharmin, M. Sivertsvik, S. S. Ray, J. Keränen and M. N. Emmambux, Development and Characterization of Poly(butylene succinate-co-adipate)/Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) with Cowpea Lignocellulosic Fibers as a Filler via Injection Molding and Extrusion Film-Casting, Macromol. Mater. Eng., 2024, 309(8), 2400037, DOI:10.1002/mame.202400037.
- D. Nath, A. K. Pal, M. Misra and A. K. Mohanty, Biodegradable Blown Film Composites from Bioplastic and Talc: Effect of Uniaxial Stretching on Mechanical and Barrier Properties, Macromol. Mater. Eng., 2023, 308(12), 2300214, DOI:10.1002/mame.202300214.
- F. Ma, B. Wang, X. Leng, Y. Wang, Z. Sun, P. Wang, L. Sang and Z. Wei, Biodegradable PBAT/PLA/CaCO3 Blowing Films with Enhanced Mechanical and Barrier Properties: Investigation of Size and Content of CaCO3 Particles, Macromol. Mater. Eng., 2022, 307(9), 2200135, DOI:10.1002/mame.202200135.
- E. d. C. D. Nunes, A. G. d. Souza, R. D. S. Coiado, E. Moura and D. d. S. Rosa, Evaluation of the Poly (Lactic Acid) and calcium carbonate effects on the mechanical and morphological properties in PBAT blends and composites, International Journal of Innovative Science, Engineering & Technology, 2017, 4(6), 2348–7968 Search PubMed.
- M. A. Azman, M. R. M. Asyraf, A. Khalina, M. Petrů, C. M. Ruzaidi, S. M. Sapuan, W. B. Wan Nik, M. R. Ishak, R. A. Ilyas and M. J. Suriani, Natural Fiber Reinforced Composite Material for Product Design: A Short Review, Polymers, 2021, 13(12), 1917, DOI:10.3390/polym13121917.
- N. K. Faheed, Advantages of natural fiber composites for biomedical applications: a review of recent advances, Emergent Mater., 2024, 7(1), 63–75, DOI:10.1007/s42247-023-00620-x.
- W. Post, L. J. Kuijpers, M. Zijlstra, M. van der Zee and K. Molenveld, Effect of Mineral Fillers on the Mechanical Properties of Commercially Available Biodegradable Polymers, Polymers, 2021, 13(3), 394, DOI:10.3390/polym13030394.
- S. Marecik, I. Pudełko-Prażuch, M. Balasubramanian, S. M. Ganesan, S. Chatterjee, K. Pielichowska, R. Kandaswamy and E. Pamuła, Effect of the Addition of Inorganic Fillers on the Properties of Degradable Polymeric Blends for Bone Tissue Engineering, Molecules, 2024, 29(16), 3826, DOI:10.3390/molecules29163826.
- H. P. S. Abdul Khalil, E. W. N. Chong, F. A. T. Owolabi, M. Asniza, Y. Y. Tye, S. Rizal, M. R. Nurul Fazita, M. K. Mohamad Haafiz, Z. Nurmiati and M. T. Paridah, Enhancement of basic properties of polysaccharide-based composites with organic and inorganic fillers: A review, J. Appl. Polym. Sci., 2019, 136(12), 47251, DOI:10.1002/app.47251.
- A. Ashfaq, N. Khursheed, S. Fatima, Z. Anjum and K. Younis, Application of nanotechnology in food packaging: Pros and Cons, J. Agric. Food Res., 2022, 7, 100270, DOI:10.1016/j.jafr.2022.100270.
- A. Hiremath, A. A. Murthy, S. Thipperudrappa and K. N. Bharath, Nanoparticles Filled Polymer Nanocomposites: A Technological Review, Cogent Eng., 2021, 8(1), 1991229, DOI:10.1080/23311916.2021.1991229.
- S. Venkatarajan and A. Athijayamani, An overview on natural cellulose fiber reinforced polymer composites, Mater. Today: Proc., 2021, 37, 3620–3624, DOI:10.1016/j.matpr.2020.09.773.
- K. Olonisakin, R. Li, X.-X. Zhang, F. Xiao, J. Gao and W. Yang, Effect of TDI-Assisted Hydrophobic Surface Modification of Microcrystalline Cellulose on the Tensile Fracture of MCC/PLA Composite, and Estimation of the Degree of Substitution by Linear Regression, Langmuir, 2021, 37(2), 793–801, DOI:10.1021/acs.langmuir.0c03130.
- C. H. Wibowo, D. Ariawan, E. Surojo and S. Sunardi, Microcrystalline Cellulose as Composite Reinforcement: Assessment and Future Prospects, Mater. Sci. Forum, 2024, 1122, 65–80, DOI:10.4028/p-viYb6d.
- E. C. Ramires, J. D. Megiatto, A. Dufresne and E. Frollini, Cellulose Nanocrystals versus Microcrystalline Cellulose as Reinforcement of Lignopolyurethane Matrix, Fibers, 2020, 8(4), 21, DOI:10.3390/fib8040021.
- M. Wang, C. He, X. Yang, G. Duan and W. Wang, Preparation and properties of PLA/PBAT composites modified with different filler particles, Mater. Lett., 2024, 372, 136960, DOI:10.1016/j.matlet.2024.136960.
- H. Peidayesh, L. Ondriš, S. Saparová, M. Kovaľaková, O. Fričová and I. Chodák, Biodegradable Nanocomposites Based on Blends of Poly(Butylene Adipate–Co–Terephthalate) (PBAT) and Thermoplastic Starch Filled with Montmorillonite (MMT): Physico-Mechanical Properties, Materials, 2024, 17(3), 540, DOI:10.3390/ma17030540.
- H. Khonakdar, S. S. Khasraghi, A. H. Yazdanbakhsh, S. R. Mousavi, S. Ahmadi, H. Arabi, M. A. L. Nobre and H. A. Khonakdar, An assessment of the role of nanosilica in thermal/thermo-oxidative degradation mechanism of poly(lactic acid)/polybutylene adipate terephthalate blend nanocomposites, Polym. Adv. Technol., 2024, 35(4), e6374, DOI:10.1002/pat.6374.
- C. Yutong and L. Yana, Development of multilayer films based on PLA/PBAT and sodium alginate for active packaging, J. Food Saf., 2024, 44(2), e13121, DOI:10.1111/jfs.13121.
- M. C. Pozza Junior, A. G. Rosenberger, F. F. da Silva, D. C. Dragunski, E. C. Muniz and J. Caetano, Application of a PLA/PBAT/Graphite sensor obtained by electrospinning on determination of 2,4,6-trichlorophenol, Environ. Technol., 2024, 45(12), 2388–2401, DOI:10.1080/09593330.2023.2173088.
- E. G. R. dos Anjos, T. R. Brazil, L. S. Montagna, G. F. de Melo Morgado, E. F. Martins, L. A. Pessan, F. K. V. Moreira, J. Marini and F. R. Passador, Biodegradation Behavior and Life Cycle Assessment of PLA/PHBV/Carbonaceous Materials Hybrid Nanocomposites for Antimicrobial Multifunctional Packaging, J. Polym. Environ., 2024, 32, 5098–5114, DOI:10.1007/s10924-024-03286-6.
- A. Surendren, A. K. Pal, A. Rodriguez-Uribe, S. Shankar, L.-T. Lim, A. K. Mohanty and M. Misra, Upcycling of post-industrial starch-based thermoplastics and their talc-filled sustainable biocomposites for single-use plastic alternative, Int. J. Biol. Macromol., 2023, 253, 126751, DOI:10.1016/j.ijbiomac.2023.126751.
- Y. Zhu, J. Fu, Z. Ou, X. Tao, T. Yan and Z. Pan, Poly(3-hydroxybutyrate)/Poly(lactic acid) Composite Nanofibrous Membrane with Outstanding Photochromic and Antibacterial Performances for Multifunctional Applications, ACS Appl. Polym. Mater., 2024, 6(16), 9883–9891, DOI:10.1021/acsapm.4c01738.
- X.-Y. Gu, L.-M. Hu, Z.-A. Fu, H.-T. Wang and Y.-J. Li, Reactive TiO2 Nanoparticles Compatibilized PLLA/PBSU Blends: Fully Biodegradable Polymer Composites with Improved Physical, Antibacterial and Degradable Properties, Chin. J. Polym. Sci., 2021, 39(12), 1645–1656, DOI:10.1007/s10118-021-2632-x.
- N. Bumbudsanpharoke, R. P. Nurhadi, B. Chongcharoenyanon, S. Kwon, N. Harnkarnsujarit and S. Ko, Effect of migration on the functionality of zinc oxide nanoparticle in polybutylene adipate terephthalate/thermoplastic starch films: A food simulant study, Int. J. Biol. Macromol., 2024, 263, 130232, DOI:10.1016/j.ijbiomac.2024.130232.
- N. Nomadolo, O. E. Dada, A. Swanepoel, T. Mokhena and S. Muniyasamy, A Comparative Study on the Aerobic Biodegradation of the Biopolymer Blends of Poly(butylene succinate), Poly(butylene adipate terephthalate) and Poly(lactic acid), Polymers, 2022, 14(9), 1894, DOI:10.3390/polym14091894.
- X. Quecholac-Piña, M. D. Hernández-Berriel, M. D. Mañón-Salas, R. M. Espinosa-Valdemar and A. Vázquez-Morillas, Degradation of Plastics under Anaerobic Conditions: A Short Review, Polymers, 2020, 12(1), 109, DOI:10.3390/polym12010109.
- J. R. Kim, J.-R. Thelusmond, V. C. Albright and Y. Chai, Exploring structure-activity relationships for polymer biodegradability by microorganisms, Sci. Total Environ., 2023, 890, 164338, DOI:10.1016/j.scitotenv.2023.164338.
- M. Brebu, Environmental Degradation of Plastic Composites with Natural Fillers—A Review, Polymers, 2020, 12(1), 166, DOI:10.3390/polym12010166.
- A. Delacuvellerie, A. Brusselman, V. Cyriaque, S. Benali, S. Moins, J.-M. Raquez, S. Gobert and R. Wattiez, Long-term immersion of compostable plastics in marine aquarium: Microbial biofilm evolution and polymer degradation, Mar. Pollut. Bull., 2023, 189, 114711, DOI:10.1016/j.marpolbul.2023.114711.
- Y.-X. Weng, Y.-J. Jin, Q.-Y. Meng, L. Wang, M. Zhang and Y.-Z. Wang, Biodegradation behavior of poly(butylene adipate-co-terephthalate) (PBAT), poly(lactic acid) (PLA), and their blend under soil conditions, Polym. Test., 2013, 32(5), 918–926, DOI:10.1016/j.polymertesting.2013.05.001.
- A. a. Zandi, A. Zanganeh, F. Hemmati and J. Mohammadi-Roshandeh, Thermal and biodegradation properties of poly(lactic acid)/rice straw composites: effects of modified pulping
products, Iran. Polym. J., 2019, 28(5), 403–415, DOI:10.1007/s13726-019-00709-3.
- G. Kurup, M. F. F. B. M. Fadzillah, N. R. R. Royan, N. A. M. Radzuan and A. B. Sulong, Impact of processing parameters on the compatibility and performance of PLA/tapioca starch biocomposites for short-term food packaging applications, Mater. Today Commun., 2025, 43, 111651, DOI:10.1016/j.mtcomm.2025.111651.
- B. Palai, S. Mohanty and S. K. Nayak, A Comparison on Biodegradation Behaviour of Polylactic Acid (PLA) Based Blown Films by Incorporating Thermoplasticized Starch (TPS) and Poly (Butylene Succinate-co-Adipate) (PBSA) Biopolymer in Soil, J. Polym. Environ., 2021, 29(9), 2772–2788, DOI:10.1007/s10924-021-02055-z.
- P. C. Mayekar, W. Limsukon, A. Bher and R. Auras, Breaking It Down: How Thermoplastic Starch Enhances Poly(lactic acid) Biodegradation in Compost—A Comparative Analysis of Reactive Blends, ACS Sustainable Chem. Eng., 2023, 11(26), 9729–9737, DOI:10.1021/acssuschemeng.3c01676.
- C. Yokesahachart, R. Yoksan, N. Khanoonkon, A. K. Mohanty and M. Misra, Effect of jute fibers on morphological characteristics and properties of thermoplastic starch/biodegradable polyester blend, Cellulose, 2021, 28(9), 5513–5530, DOI:10.1007/s10570-021-03921-8.
- E.-R. Radu, D. M. Panaitescu, C.-A. Nicolae, R. A. Gabor, V. Rădiţoiu, S. Stoian, E. Alexandrescu, R. Fierăscu and I. Chiulan, The Soil Biodegradability of Structured Composites Based on Cellulose Cardboard and Blends of Polylactic Acid and Polyhydroxybutyrate, J. Polym. Environ., 2021, 29(7), 2310–2320, DOI:10.1007/s10924-020-02017-x.
- M. Dammak, Y. Fourati, Q. Tarrés, M. Delgado-Aguilar, P. Mutjé and S. Boufi, Blends of PBAT with plasticized starch for packaging applications: Mechanical properties, rheological behaviour and biodegradability, Ind. Crops Prod., 2020, 144, 112061, DOI:10.1016/j.indcrop.2019.112061.
- M. Danko, K. Mosnáčková, A. Vykydalová, A. Kleinová, A. Puškárová, D. Pangallo, M. Bujdoš and J. Mosnáček, Properties and Degradation Performances of Biodegradable Poly(lactic acid)/Poly(3-hydroxybutyrate) Blends and Keratin Composites, Polymers, 2021, 13(16), 2693, DOI:10.3390/polym13162693.
- R. Das, E. J. Curry, T. T. Le, G. Awale, Y. Liu, S. Li, J. Contreras, C. Bednarz, J. Millender, X. Xin, D. Rowe, S. Emadi, K. W. H. Lo and T. D. Nguyen, Biodegradable nanofiber bone-tissue scaffold as remotely-controlled and self-powering electrical stimulator, Nano Energy, 2020, 76, 105028, DOI:10.1016/j.nanoen.2020.105028.
- L. Cui, J. Zhang, J. Zou, X. Yang, H. Guo, H. Tian, P. Zhang, Y. Wang, N. Zhang, X. Zhuang, Z. Li, J. Ding and X. Chen, Electroactive composite scaffold with locally expressed osteoinductive factor for synergistic bone repair upon electrical stimulation, Biomaterials, 2020, 230, 119617, DOI:10.1016/j.biomaterials.2019.119617.
- R. Borah, J. Upadhyay and K. Acharjya, Functionalized polyaniline:chitosan nanocomposites as a potential biomaterial, Mater. Today: Proc., 2020, 32, 334–343, DOI:10.1016/j.matpr.2020.01.583.
- A. Vlachopoulos, G. Karlioti, E. Balla, V. Daniilidis, T. Kalamas, M. Stefanidou, N. D. Bikiaris, E. Christodoulou, I. Koumentakou, E. Karavas and D. N. Bikiaris, Poly(Lactic Acid)-Based Microparticles for Drug Delivery Applications: An Overview of Recent Advances, Pharmaceutics, 2022, 14(2), 359, DOI:10.3390/pharmaceutics14020359.
- Y. Jiaying, S. Bo, W. Xiaolu, Z. Yanyan, W. Hongjie, S. Nan, G. Bo, W. Linna, Z. Yan, G. Wenya, L. Keke, J. Shan, L. Chuan, Z. Yu, Z. Qinghe and Z. Haiyu, Arenobufagin-loaded PEG-PLA nanoparticles for reducing toxicity and enhancing cancer therapy, Drug Delivery, 2023, 30(1), 2177362, DOI:10.1080/10717544.2023.2177362.
- Y. Jiaying, S. Bo, W. Xiaolu, Z. Yanyan, W. Hongjie, S. Nan, G. Bo, W. Linna, Z. Yan, G. Wenya, L. Keke, J. Shan, L. Chuan, Z. Yu, Z. Qinghe and Z. Haiyu, Arenobufagin-loaded PEG-PLA nanoparticles for reducing toxicity and enhancing cancer therapy, Drug Delivery, 2023, 30(1), 2177362, DOI:10.1080/10717544.2023.2177362.
- D. Loca, E. Sevostjanovs, M. Makrecka, O. Zharkova-Malkova, L. Berzina-Cimdina, V. Tupureina and M. Sokolova, Microencapsulation of mildronate in biodegradable and non-biodegradable polymers, J. Microencapsulation, 2014, 31(3), 246–253, DOI:10.3109/02652048.2013.834992.
- C. Chen, L. Wang, S. Shams Es-haghi, M. Tajvidi, J. Wang and D. J. Gardner, Biodegradable and recyclable bio-based laminated films of poly (lactic acid) and cellulose nanocrystals for food barrier packaging, Food Packag. Shelf Life, 2024, 42, 101244, DOI:10.1016/j.fpsl.2024.101244.
- U. Lawal, R. Samyuktha, V. Robert, K. Sreelakshmi, A. Gopi, M. Poochi, S. Loganathan, S. Thomas and R. B. Valapa, Poly(lactic acid)/cholecalciferol based composites for active food packaging application, Int. J. Biol. Macromol., 2023, 246, 125637, DOI:10.1016/j.ijbiomac.2023.125637.
- M. Cvek, U. C. Paul, J. Zia, G. Mancini, V. Sedlarik and A. Athanassiou, Biodegradable Films of PLA/PPC and Curcumin as Packaging Materials and Smart Indicators of Food Spoilage, ACS Appl. Mater. Interfaces, 2022, 14(12), 14654–14667, DOI:10.1021/acsami.2c02181.
- K. Olonisakin, A. Wen, S. He, H. Lin, W. Tao, S. Chen, W. Lin, R. Li, X.-x. Zhang and W. Yang, The Development of Biodegradable PBAT-Lignin-Tannic Acid Composite Film: Properties, Biodegradability, and Potential Barrier Application in Food Packaging, Food Bioprocess Technol., 2023, 16(7), 1525–1540, DOI:10.1007/s11947-023-02997-3.
- X. Diao, C. Zhang and Y. Weng, Properties and Degradability of Poly(Butylene Adipate-Co-Terephthalate)/Calcium Carbonate Films Modified by Polyethylene Glycol, Polymers, 2022, 14(3), 484, DOI:10.3390/polym14030484.
- A. G. Souza, R. R. Ferreira, L. C. Paula, S. K. Mitra and D. S. Rosa, Starch-based films enriched with nanocellulose-stabilized Pickering emulsions containing different essential oils for possible applications in food packaging, Food Packag. Shelf Life, 2021, 27, 100615, DOI:10.1016/j.fpsl.2020.100615.
- R.-M. Wang, S.-R. Zheng and Y.-P. Zheng, Introduction to polymer matrix composites, Polymer Matrix Composites and Technology, 2011, 1–548, DOI:10.1533/9780857092229.1.
- H. Oliver-Ortega, F. Julián, F. X. Espinach and J. A. Méndez, Simulated Environmental Conditioning of PHB Composites Reinforced with Barley Fibres to Determine the Viability of Their Use as Plastics for the Agriculture Sector, Polymers, 2023, 15(3), 579, DOI:10.3390/polym15030579.
- C. Maraveas, The Sustainability of Plastic Nets in Agriculture, Sustainability, 2020, 12(9), 3625, DOI:10.3390/su12093625.
- C. Peñalva, M. Pérez, F. Braca and D. Redondo, Reducing the Effects of Plastic Waste in Agricultural Applications by Developing New Ok Soil Biodegradable Plastics, Detritus, 2020,(13), 67–77, DOI:10.31025/2611-4135/2020.14023.
- K. Rajpoot, N. Desai, H. Koppisetti, M. Tekade, M. C. Sharma, S. K. Behera and R. K. Tekade, Chapter 14 - In silico methods for the prediction of drug toxicity, in Pharmacokinetics and Toxicokinetic Considerations, ed. R. K. Tekade, Academic Press, 2022, pp. 357–383. DOI:10.1016/B978-0-323-98367-9.00012-3.
- L. Zimmermann, Z. Bartosova, K. Braun, J. Oehlmann, C. Völker and M. Wagner, Plastic Products Leach Chemicals That Induce In Vitro Toxicity under Realistic Use Conditions, Environ. Sci. Technol., 2021, 55(17), 11814–11823, DOI:10.1021/acs.est.1c01103.
- S. Ali, Isha and Y.-C. Chang, Ecotoxicological Impact of Bioplastics Biodegradation: A Comprehensive Review, Processes, 2023, 11(12), 3445, DOI:10.3390/pr11123445.
- L. Filiciotto and G. Rothenberg, Biodegradable Plastics: Standards, Policies, and Impacts, ChemSusChem, 2021, 14(1), 56–72, DOI:10.1002/cssc.202002044.
- A. D5152-91. Practice for Water Extraction of Residual Solids from Degraded Plastics for Toxicity Testing (Withdrawn 1998). 1991; available from: https://www.astm.org/d5152-91.html.
- A. D5951-96(2002). Standard Practice for Preparing Residual Solids Obtained After Biodegradability Standard Methods for Plastics in Solid Waste for Toxicity and Compost Quality Testing (Withdrawn 2011). 1996; available from: https://www.astm.org/d5951-96r02.html.
- P. A. Palsikowski, M. M. Roberto, L. R. D. Sommaggio, P. M. S. Souza, A. R. Morales and M. A. Marin-Morales, Ecotoxicity Evaluation of the Biodegradable Polymers PLA, PBAT and its Blends Using Allium cepa as Test Organism, J. Polym. Environ., 2018, 26(3), 938–945, DOI:10.1007/s10924-017-0990-9.
- S. Sforzini, L. Oliveri, S. Chinaglia and A. Viarengo, Application of Biotests for the Determination of Soil Ecotoxicity after Exposure to Biodegradable Plastics, Frontiers in Environmental Science, 2016, 4(68) DOI:10.3389/fenvs.2016.00068.
- F. Farachi, G. Bettas Ardisson and F. Degli Innocenti, Handbook of Biodegradable Polymers 3. Environmental fate and ecotoxicity assessment of biodegradable polymers, ed. C. Bastioli, De Gruyter, 2020, pp. 45–74. DOI:10.1515/9781501511967-003.
- C. Catarci Carteny, E. D. Amato, F. Pfeiffer, C. Christia, N. Estoppey, G. Poma, A. Covaci and R. Blust, Accumulation and release of organic pollutants by conventional and biodegradable microplastics in the marine environment, Environ. Sci. Pollut. Res. Int., 2023, 30(31), 77819–77829, DOI:10.1007/s11356-023-27887-1.
- A. Hejna, Renewable, Degradable, and Recyclable Polymer Composites, Polymers, 2023, 15(7), 1769, DOI:10.3390/polym15071769.
- K. L. Ang, E. T. Saw, W. He, X. Dong and S. Ramakrishna, Sustainability framework for pharmaceutical manufacturing (PM): A review of research landscape and implementation barriers for circular economy transition, J. Cleaner Prod., 2021, 280, 124264, DOI:10.1016/j.jclepro.2020.124264.
- J. H. Song, R. J. Murphy, R. Narayan and G. B. Davies, Biodegradable and compostable alternatives to conventional plastics, Philos. Trans. R. Soc., B, 2009, 364(1526), 2127–2139, DOI:10.1098/rstb.2008.0289.
- M. S. Ayilara, O. S. Olanrewaju, O. O. Babalola and O. Odeyemi, Waste Management through Composting: Challenges and Potentials, Sustainability, 2020, 12(11), 4456, DOI:10.3390/su12114456.
- M. Yang, L. Chen, J. Wang, G. Msigwa, A. I. Osman, S. Fawzy, D. W. Rooney and P.-S. Yap, Circular economy strategies for combating climate change and other environmental issues, Environ. Chem. Lett., 2023, 21(1), 55–80, DOI:10.1007/s10311-022-01499-6.
- T. D. Paul Ekins, P. Drummond, R. Bleischwitz, N. Hughes and L. Lotti, The Circular Economy: What, Why, How and Where”, Background paper for an OECD/EC Workshop on 5 July 2019 within the workshop series “Managing environmental and energy transitions for regions and cities”. 2019.
- S. Kara, M. Hauschild, J. Sutherland and T. McAloone, Closed-loop systems to circular economy: A pathway to environmental sustainability?, CIRP Ann., 2022, 71(2), 505–528, DOI:10.1016/j.cirp.2022.05.008.
- J. Hopewell, R. Dvorak and E. Kosior, Plastics recycling: challenges and opportunities, Philos. Trans. R. Soc., B, 2009, 364(1526), 2115–2126, DOI:10.1098/rstb.2008.0311.
- H. Winkler, Closed-loop production systems—A sustainable supply chain approach, CIRP J. Manuf. Sci. Technol., 2011, 4(3), 243–246, DOI:10.1016/j.cirpj.2011.05.001.
- N. M. P. Bocken, L. Niessen and S. W. Short, The Sufficiency-Based Circular Economy—An Analysis of 150 Companies, Frontiers in Sustainability, 2022, 3, 899289, DOI:10.3389/frsus.2022.899289.
- O. Platnieks, A. Barkane, N. Ijudina, G. Gaidukova, V. K. Thakur and S. Gaidukovs, Sustainable tetra pak recycled cellulose/Poly(Butylene succinate) based woody-like composites for a circular economy, J. Cleaner Prod., 2020, 270, 122321, DOI:10.1016/j.jclepro.2020.122321.
- P. McKeown and M. D. Jones, The Chemical Recycling of PLA: A Review, Sustainable Chem., 2020, 1, 1–22, DOI:10.3390/suschem1010001.
- G. Gaidukova, O. Platnieks, A. Aunins, A. Barkane, C. Ingrao and S. Gaidukovs, Spent coffee waste as a renewable source for the production of sustainable poly(butylene succinate) biocomposites from a circular economy perspective, RSC Adv., 2021, 11(30), 18580–18589, 10.1039/d1ra03203h.
- N. A. Tarazona, R. Machatschek, J. Balcucho, J. L. Castro-Mayorga, J. F. Saldarriaga and A. Lendlein, Opportunities and challenges for integrating the development of sustainable polymer materials within an international circular (bio)economy concept, MRS Energy Sustainability, 2022, 9(1), 28–34, DOI:10.1557/s43581-021-00015-7.
- L. G. Pinaeva and A. S. Noskov, Biodegradable biopolymers: Real impact to environment pollution, Sci. Total Environ., 2024, 947, 174445, DOI:10.1016/j.scitotenv.2024.174445.
- V. Titone, L. Botta and F. P. La Mantia, Mechanical Recycling of New and Challenging Polymer Systems: A Brief Overview, Macromol. Mater. Eng., 2025, 310(1), 2400275, DOI:10.1002/mame.202400275.
- K. Ragaert, L. Delva and K. Van Geem, Mechanical and chemical recycling of solid plastic waste, Waste Manage., 2017, 69, 24–58, DOI:10.1016/j.wasman.2017.07.044.
- J. Kirchherr, N.-H. N. Yang, F. Schulze-Spüntrup, M. J. Heerink and K. Hartley, Conceptualizing the Circular Economy (Revisited): An Analysis of 221 Definitions, Resour., Conserv. Recycl., 2023, 194, 107001, DOI:10.1016/j.resconrec.2023.107001.
- A. P. M. Velenturf and P. Purnell, Principles for a sustainable circular economy, Sustainable Prod. Consum., 2021, 27, 1437–1457, DOI:10.1016/j.spc.2021.02.018.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.