You are on page 1of 20

Poly(butylene-succinate)-based blends with enhanced oxygen permeability

Nina Kantor- Malujdy1, Sandra Skowron1,#, Beata Michalkiewicz2,


Miroslawa El Fray1*
1Department of Polymer and Biomaterials Science,
Faculty of Chemical Technology and Engineering,
West Pomeranian University of Technology in Szczecin
Al. Piastow 45, 70-311 Szczecin, Poland
2 Department of Catalytic and Sorbent Materials Engineering,
Faculty of Chemical Technology and Engineering,
West Pomeranian University of Technology in Szczecin
Pułaskiego 10, 70-322 Szczecin, Poland

*Corresponding author: mirfray@zut.edu.pl

# current address: Azoty Polyolefins Group, Police, ul. Kuznicka 1, 72-010 Police,
POLAND
Abstract

Poly(butylene succinate) (PBS) is a biodegradable polymer produced from renewable raw

materials and is widely used in the production of packaging materials. Among important

1
features, oxygen permeability is crucial for “breathing” packaging materials (such as

packaging for medical devices sterilized with gaseous ethylene oxide). Recently,

poly(butylene succinate-dilinolene succinate) (PBS-DLS) copolymers containing long

chains of fatty acids and showing excellent elasticity were synthesized. In this work, bio-

based polymer blends of PBS and PBS-DLS copolymer with an aliphatic-aromatic

poly(butylene terephthalate-butylene adipate) (PBAT) (Ecoflex) were prepared in order

to further improve their oxygen permeability while maintaining mechanical stability. PBS

and PBS-DLS copolymer containing 90 wt% of hard segments and 10 wt% of soft

segments, respectively, were used for blends preparation with 10 wt% of PBAT. The

chemical structure was analyzed using infrared spectroscopy and thermal properties were

determined with differential scanning calorimetry. The introduction of PBAT to the

blends did not affect their melting temperatures. PBS-based blends were miscible at the

molecular level based on single Tg value calculated from Fox equation and uniform

fracture surface morphology from scanning electron microscopy (SEM). Importantly, the

oxygen permeability of PBS-DLS/PBAT blend was comparable to non-biodegradable

high-density polyethylene (HDPE) Tyvek used for medical devices packaging.

Key words: polyester blends, biodegradable polymers, polybutylene succinate, oxygen

permeability

1. Introduction

Serious economic and environmental problems are the result of ineffective waste

management and the use of petroleum-based plastics such as polyolefins (polyethylene,

polypropylene). Economic growth in the 21 st century has contributed to an increase in

plastic consumption and thus pollution 1. In addition, the SARS-CoV-2 virus pandemic

2
caused an even greater increase in waste production by increased use of disposable
2,3.
personal protective equipment For this reason, there is a pressing need to introduce
4–7.
biodegradable materials for the fabrication of medical devices or their packaging

Currently, aliphatic polyesters are considered to be one of the most promising candidates

as they are susceptible to biodegradation 8. They are recyclable (chemically or

biologically) without a negative impact on the environment, along with their excellent
9,10.
properties For this reason, some thermoplastics have already been replaced with

biodegradable and bio-based polyesters based on succinic acid for the production of
11,12.
packaging materials Poly(butylene succinate) (PBS) has good barrier properties,

which collectively depend on the polarity of the polymer, chain symmetry, crystallinity,
13 .
molecular weight, orientation and temperature However, PBS has poor water

permeability and elasticity. To enhance PBS elasticity and reduce crystallinity, a


14
dilinoleic acid or dilinoleic diol derived from plant oil has recently been used for the

synthesis of poly[(butylene succinate)-co-(dilinoleic succinate)] copolyesters (PBS-DLS)

by two-step synthesis using titanium oxide/silicon dioxide (TiO2/SiO2) catalyst via melt

polycondensation15. The polymers are semi-crystalline and the butylene succinate

segments characterized by high crystallinity are responsible for the stiffness of the
13,16
materials while the soft segments containing amorphous dilinoleic succinate (DLS)
14.
impart polymers' flexibility, depending on their weight fraction In this way, it is

possible to obtain materials with various properties, ranging from rigid to rubber-like,

which allows the properties to be tailored to a specific application. The possibility of

tailoring PBS-DLS copolymers properties make them useful for various applications,

including coatings for fertilizers 17, drugs encapsulation 18 or tissue engineering 7,15.

3
Poly(butylene terephthalate-butylene adipate) (PBAT), known under Ecoflex brand

name, is being currently used as an alternative to non-biodegradable plastics 16,19 . PBAT

is considered a semi-aromatic, thermoplastic copolyester. It is obtained from

petrochemical raw materials, but it is almost 100% biodegradable, and its degradation

products are non-toxic. PBAT has a matrix-domain structure and it is made of 20-30 nm

crystal domains formed by aromatic PBT sequences arranged in an amorphous matrix.

The commercial blend of PLA and PBAT known as Ecovio is a promising polymeric

material for packaging applications20–23. PBS/PBAT blends are used in the packaging

industry and have similar mechanical properties (resilience and flexibility24) to films

made of low-density polyethylene. This type of blend can be obtained with the use of

extrusion technology 11,25. In this work, we analyze the effect of PBAT (Ecoflex) on the

oxygen permeability of new blends based on PBS-DLS and PBAT. Likewise, the thermal

and mechanical analyses were performed to demonstrate the potential of new blends in

the packaging applications.

2. Materials and Methods

2.1.Materials

Poly[(butylene succinate)-co-(dilinoleic succinate)] (PBS-DLS) copolyesters with 90:10

wt% hard to soft segments ratio were synthesized with the use of heterogeneous titanium

dioxide/silicon dioxide (TiO2/SiO2) catalyst according to the procedure previously


15
described in . Briefly, the first step of synthesis was an esterification reaction. 1,4-

butanediol (BD, Merck Germany) and succinic acid (SA, Alfa Aestar, Germany) were

copolymerized at 180°C in the presence of TiO2/SiO2 catalyst (Hunstman, Germany).

After collecting 95% of the theoretical amount of water, bio-based dilinoleic diol

(PripolTM 2033, Croda, The Netherlands) and a second dose of TiO2/SiO2 catalyst were

4
added to the reaction mixture. Polycondensation started when the pressure was decreased

to 0.2-0.4 hPa and the temperature increased to 240°C. Bio-based poly(butylene

succinate) (PBS) (BioPBS FZ79AC, ptt MCC, Biochem, Germany) and poly(butylene

terephthalate-butylene adipate) (PBAT) (Ecoflex F BX 7011, BASF) were used as

received. PBS and PBS-DLS blends with Ecoflex in 90/10 wt.% ratio were prepared with

a "ZAMAK MERCATOR" twin-screw extruder with L/D=22. The processing

temperature was set to different temperatures at different zones: 80 °C in the 1st zone, 100

°C in the 2nd zone, 120 °C in the 3rd zone, and 140 ̊C at the extrusion head. The process

was carried out at a speed of 100 rpm/min. The polymer granules were dosed with a

customized pneumatic feeder at a pressure of 5 bar. A pneumatic dispenser was used due

to the large diameter of the granules. Polymer blends were extruded through the nozzle

of 1 mm diameter, and the materials were cooled in a water bath and next granulated.

2.2.Methods

The chemical structure was evaluated using an Alpha Bruker FT-IR spectrometer with an

ATR attachment (Bruker Optik, Ettlingen, Germany) in the wavenumber range 400-4000

cm-1 with an accuracy of 2 cm-1. 32 scans were made for each of the tested materials and

the sample measurement time was approximately 70 seconds. The thermal properties

were examined using a TA Instruments Discovery 2500 (TA Instruments Inc., New

Castle, DE, USA) differential scanning calorimeter (DSC). The heating and cooling rate

was 10°C/min and the measurements were carried out under N2 atmosphere. The research

̊ to 200 ̊C; cooling


procedure was based on the use of three cycles: first heating from -90 C

of test samples from 200 ̊C to -90 ̊C; second heating from -90 C
̊ to 200 ̊C. The glass

transition (Tg), crystallization (Tc) and melting (Tm) temperatures were determined using

Trios software.

5
Ultra high-resolution scanning electron microscopy (UHR SEM, SU82020, Hitachi,

Tokyo, Japan) was used to examine the blend morphology. The specimens were fractured

in liquid nitrogen, and the fracture surface was examined. All the specimens were coated

with a thin 10 nm chromium layer prior to SEM analysis, in order to avoid electrostatic

charging. Micrographs were taken in high-vacuum mode, at an acceleration voltage of

15-20 kV.

The melt flow index (MFI) was measured based on the ISO 1133 standard (PN-93/C-

89069) using a CAEST type CA-MAN-001 plastometer and the applied load was 2.16

kg. The cut-off time was 30 s. The MFI was calculated according to the formula: MFI =

600 × mt , where 600 stands for the reference time (10 min = 600 s), m stands for the

mass of the sample after the desired period (g), and t stands for the period of time (s).

Mechanical properties were examined following the PN-ISO 527-3/5 standard. Initially,

evaluated materials were processed into foils using the “Remi-plast” hydraulic hot press

with the 6-ton pressure and at the temperature of 20°C degrees higher than the melting

point of individual materials/blend components. A frame with dimensions of 10 cm x 10

cm and a thickness of 0.5 mm was used to control samples geometry. Testing samples,

pounched with hydraulic pouncher (Endeavour Lab, Poland) into a dog-bone shape of

type 3 were conditioned for 24 hours at a temperature of 22 ̊C prior to mechanical tests.

Tensile properties were assessed with the use of an Instron 3366 machine with a 500N

load cell. Measurements were made with a testing speed of 100 mm/min. The tests were

carried out at a temperature of 22 °C with 50% of humidity. The hardness Shore D was

determined according to the PN-ISO 868 standard using 4 mm thick samples. Finally, the

oxygen transmission rate was tested using the Mocon OX-TRAN 2/20 device, based on

the ASTM F 1927-14 standard, at a temperature of 22.5°C and a humidity of 0%. Before

6
starting the measurement of the oxygen transmission rate, samples were prepared in the

form of very thin, homogeneous films using the “Remi-plast” hydraulic hot press with

the 6-ton pressure. Each sample was pressed twice to obtain the most homogeneous

bubble-free material possible. Materials in the form of films with a thickness from 67 μm

to 83 μm were obtained.

3. Results and Discussion

3.1.Infrared Spectroscopy (ATR-FTIR)

To assess the chemical structure of prepared blends, characteristic bands were recorded

with ATR-FTIR spectrophotometer (Figure 1). PBS and PBS-DLS blends with Ecoflex

revealed the increased intensity of the band characteristic for the presence of methylene

groups (-CH2-) in the 2945 - 2922 cm-1 range and the carbonyl moiety (C=O) in the 1712-

1713 cm-1. Subsequently, the bands at 1100-1200 cm-1characteristic for the ester moiety

(C-O) were observed along with the band at 600-700 cm-1 typical for alkyl moiety (C-

H). The intense band at 1504 cm-1 has been observed, which is characteristic for stretching

peaks of the C=C phenylene group. It is assumed that phthalate units present in PBAT

were incorporated into PBS/PBAT blend. Moreover, by analyzing the ATR-FTIR spectra

of PBS-DLS/PBAT blend, it can be seen that there are no chemical interactions between

the copolymer and PBAT since no compatibilizer was used for the preparation of the

blend.

7
Figure 1 ATR-FTIR spectra of PBS, PBS-DLS copolymer and their blends with 10 wt%

of PBAT

3.2.Thermal properties and melt flow index

Thermal analysis of the tested materials was performed with the use of differential

scanning calorimetry (DSC). First heating exotherms were not analysed since they are

reflecting thermal history of the materials. Based on DSC thermograms, the values of the

glass transition temperature, melting temperature and crystallization temperature are

detailed in Table 1 and showed in Figures 2 and 3. The melting temperature (Tm) of PBS

material and PBS/PBAT 90/10 blend are not changed and were found at 113.9°C.

Interestingly, the crystallization temperature (Tc) of PBS/PBAT 90/10 blend has been

shifted to higher temperatures (83.5 ̊C) in comparison to PBS material (Tc=62.2 ̊C), thus

demonstrating that the addition of PBAT accelerates crystallization and probably a new

crystalline phase formation. Presumably, aromatic moieties in PBAT are acting as

nucleating agent, thus shifting the crystallization temperature of the blend towards higher

values. In the case of PBS-DLS-based blend, the melting temperature did not change in

comparison PBS-DLS copolymer (109.0 ̊C). In contrary to PBS-based blends, the

8
crystallization temperature of PBS-DLS/PBAT 90/10 blend has shifted to lower

temperatures (from 80.9 °C to 73.1 °C). This behaviour can be explained by plasticizing

effect of DLS component containing long amorphous sequences of fatty diol.

Glass transition temperatures, Tg of a binary polymer blend can reveal whether two
26
components are miscible indicating full miscibility in case of detecting single T g .

Following Fox equation (1):

1 𝑥(𝑃𝐵𝑆) 𝑥(𝑃𝐵𝐴𝑇)
= 𝑇𝑔(𝑃𝐵𝑆) + 𝑇𝑔(𝑃𝐵𝐴𝑇) (1)
𝑇𝑔

The detected single Tg for PBS/PBAT blend (-36.6 °C) is close to the calculated one (-39

°C) thus indicating blend miscibility. The glass transition temperature found for PBS-

DLS/PBAT (-41,4 °C) is almost identical to the calculated one (-41 °C) thus confirming

that PBS and PBS-DLS are miscible with PBAT at given weigh ratio.

9
Figure 2 The DSC second heating thermograms for PBS, PBS-DLS compolymer and

their blends with 10 wt% of PBAT

10
Figure 3 The DSC cooling thermograms for PBS, PBS-DLS copolymer and their blends

with 10 wt% of PBAT

Table 1 Summary of thermal analysis and melt flow index for PBS, PBS/DLS copolymer

and their blends with 10 wt% of PBAT

ΔHm ΔCp MFI


Tm Tc ΔHc Tg
Samples
[ ̊C] [J/g] [ ̊C] [J/g] [ ̊C] [J/(g ̊C)] [g/10 min]

PBS 113.9 69.53 62.2 60.06 -40.2 0.056 15

PBS-DLS 109.0 54.39 80.9 61.68 -41.5 0.138 10

PBAT 104.6 45.86 63.4 63.41 -27.5 0.050 4

PBS/PBAT
113.9 50.82 83.5 60.33 -36.6 0.074 4
90/10

11
PBS-

DLS/PBAT 109.0 46.67 73.1 52.21 -41.4 0.096 8

90/10

Tg—glass transition temperature, ΔCp—change of heat capacity, Tc—crystallization

temperature, ΔHc—crystallization enthalpy, Tm—melting temperature, ΔHm—melting

enthalpy

The analysis of melt viscosity (Table 1) shows that the incorporation of 10 wt.% PBAT,

both into PBS and PBS-DLS-based blends, resulted in lower MFI values. Materials with

a low melt flow index are characterized by high melt viscosity and are suitable for

extrusion processing, while materials with a high melt flow index are characterized by
24.
low melt viscosity and are more suitable for injection moulding Thus, the results of

MFI of 4 and 8 g/10 min derived for blends containing 10 wt% PBAT clearly indicate

their attractiveness for film blowing or casting, the most popular processing methods used

for packaging applications.

3.3. Blends morphology from SEM analysis

The SEM micrographs of fracture surfaces (Figure 4) of polymer blends are evidencing

good miscibility of blend components. Homogenous morphology, free of inclusions,

indicating lack of strong phase separation can be seen for both PBS/PBAT (Figure 4a)

and PBS-DLS/PBAT (Figure 4b) blends. Importantly, in case of blend containing PBS-

DLS copolymer (Figure 4b) bearing biobased dilinoleic diol, more rough but homogenous

fracture surface morphology can be seen.

12
a) b)

Figure 4. Micrographs of fracture surfaces of blends. a) PBS/PBAT; b) PBS-DLS/PBAT

3.4.Mechanical properties and hardness

As shown in Table 2 and in Figure 5, the mechanical properties of PBS-based

materials and PBAT polymer prepared by press moulding thus showing isotropy are

significantly different indicating higher elasticity and lower mechanical strength of PBAT

thus revealing its ductile character. Commercially available PBS and custom synthesized

PBS-DLS copolymer show comparable values of tensile strength and modulus, however

amorphous DLS acts as plasticizer, thus decreasing slightly mechanical properties of

PBS-DLS copolymer. The incorporation of 10 wt% PBAT to PBS and PBS-DLS

copolymer contributed to the slight reduction of the modulus of elasticity the obtained

blends indicating the loss of ductility due to incorporation of PBAT. Overall, mechanical

properties of PBS/PBAT 90/10 and PBS-DLS/PBAT 90/10 blends show stress at break

of 35 and 25 MPa which is comparable to commercial grades of high denisty polyetylene

(HDPE) (30 MPa)19.

13
Figure 5 Stress at break for PBS, PBS-DLS and their blends with 10 wt% of PBAT

Table 2 Mechanical properties of PBS, PBS/DLS 90/10 and their blends with 10 wt% of

PBAT

Tensile Stress at Young’s


Yield Strain at
Samples Strength Break Modulus Hardness
Stress Break
[MPa] [MPa] [MPa] Shore D
[MPa] [%]

PBS 41.8±2.0 36.11±4.8 22.3±5 27±15 561±21 61

PBS-DLS 34.1±0.9 33.7±0.9 19.3±2 18±2 481±30 65

PBAT 10.4±1.5 10.3±0.2 5.1±2.4 162±13 94±5 35

PBS/PBAT
37.3±1.3 35.5±3.5 27.3±4.0 13±3 555±24 62
90/10

14
PBS-

DLS/PBAT 26.9±0.7 25.9±1.7 15.8±2.0 18±2 359±14 53

90/10

The components and the extruded PBS/PBAT 90/10 and PBS-DLS/PBAT 90/10 blends

show hardness of 35 to 65 Shore D with the lowest value of 35 found for PBAT material.

Overall, the prepared blends showed similar hardness values of 62-53 Shore D. These

values are also comparable to the hardness Shore D of HDPE which is in a range of 55-

67 20.

3.5.Oxygen permeability

The oxygen permeability is an important parameter to be considered for packaging

applications. However, while low oxygen permeability is needed for food packaging

(ensuring good barrier properties), the packaging of medical products (sterile instruments,

implants, biomaterials etc.) requires high permeability allowing penetration of gaseous

sterilizing agent (ethylene oxide). High oxygen permeability of HDPE known as Tyvek
20 is mainly originating from its apolar properties and microporosity of spun-bond fibres
25
. Among other biodegradable polymers, PBAT has slightly better gas barrier properties

15
20. The PBS-DLS/PBAT 90/10 blend has shown relatively high oxygen permeability

(Table 3).

Table 3. Oxygen permeability results

Sample Oxygen

Sample thickness permeability


3 2
[µm] [cm /(m x24h)]

HDPE (Tyvek) -* 500-700

PBS-DLS 105 239

PBS 70-80 400 20

PBAT 70-80 550 20

PBS/PBAT 90/10 67 80

PBS-DLS/PBAT 90/10 75 326

*- data not available

Interestingly, the oxygen permeability of PBS-DLS copolymer is also high due to the

presence of long aliphatic chains of fatty alcohol resulting in low crystallinity thus leading

to less tortuous path of oxygen penetration through the amorphous phase. Likewise, it is

contributing to high diffusion of amorphous regions and higher diffusivity of oxygen as

the permeant. The introduction of PBAT further reduced the packaging density of the

polymer chains 24 thus leading to improved permeability of PBS-DLS/PBAT 90/10 blend.

Thus, the PBS-DLS/PBAT 90/10 blend can be considered as an interesting alternative to

the Tyvek.

16
4. Conclusions

Bio-based poly(butylene succinate) (PBS) and PBS-based copolymers were blended with

an aliphatic-aromatic poly(butylene terephthalate-butylene adipate) (PBAT) (Ecoflex). It

was confirmed that melting temperature of new blends were similar to the neat blend

components. However, for PBS/PBAT 90/10 blend we noticed higher crystallization

temperature compared to the neat PBS thus demonstrating that the addition of PBAT

accelerates crystallization and probably a new crystalline phase formation. Hildebrand

solubility parameters of PBS and PBAT are close thus predicting thearetically good

miscibility. Indeed, the DSC resuts revealed single glass transition temperature for binary

polymer blends thus indicating blend miscibility. The SEM micrographs of fracture

surfaces of polymer blends have also evidenced good miscibility of blend components.

The incorporation of 10 wt% PBAT to PBS and PBS-DLS copolymer contributed to the

reduction of the modulus of elasticity the obtained blends confirming the plasticizing

effect of PBAT but maintaining the tensile strength at the values between 25 and 35 MPa

being in the range of HDPE (30 MPa). The hardness Shore D values of prepared blends

were 62 and 53, respectively, and also in the range of hardness Shore D of HDPE (55 to

67). The new copolymer PBS-DLS and its blend with PBAT turned out to be the best

material in terms of oxygen permeability. The presence of long aliphatic chains of fatty

alcohol (DLS sequences) resulted in low crystallinity and thus less tortuous path of

oxygen penetration through the amorphous phase. The introduction of PBAT further

reduced the packaging density of the polymer chains thus leading to improved

permeability of PBS-DLS/PBAT 90/10 blend. Overall, the oxygen permeability for PBS-

DLS-PBAT 90/10 blend is close to HDPE values, which makes this blend attractive for

medical devices packaging.

17
Acknowledgements

The authors acknowledge funding from the European Union's Horizon 2020 research and
innovation program under Marie Skłodowska-Curie grant agreement no. 872152
(GREEN-MAP).

5. References
1. Reddy M, Vivekanandhan S, Misra M, Bhatia S, Mohanty A Biobased plastics and
bionanocomposites: Current status and future opportunities. Progress in Polymer
Science. 2013;38(10-11):1653-1689. doi:10.1016/j.progpolymsci.2013.05.006

2. Singh N, Tang Y, Zhang Z, Zheng C. COVID-19 waste management: Effective and


successful measures in Wuhan, China. Resources, Conservation and Recycling.
2020;163(July):10-11. doi:10.1016/j.resconrec.2020.105071

3. Breese D, Hatfield E, Corning A, et al. Packaging Technology. In: The Wiley


Encyclopedia of Packaging Technology. John Wiley & Sons, Inc.; 2010:685-806.
doi:10.1002/9780470541395.ch12

4. Niaounakis M. Recycling of biopolymers – The patent perspective. European Polymer


Journal. 2019;114(February):464-475. doi:10.1016/j.eurpolymj.2019.02.027

5. Yin G, Yang X. Biodegradable polymers: a cure for the planet, but a long way to go.
Journal of Polymer Research. 2020;27(2):38. doi:10.1007/s10965-020-2004-1

6. Nikolaivits E, Pantelic B, Azeem M, et al. Progressing plastics circularity: A review of


mechano-biocatalytic approaches for waste plastic (re)valorization. Frontiers in
Bioengineering and Biotechnology. 2021;9(June):535. doi:10.3389/fbioe.2021.696040

7. Gigli M, Fabbri M, Lotti N, Gamberini R, Rimini B, Munari A. Poly(butylene


succinate)-based polyesters for biomedical applications: A review in memory of our
beloved colleague and friend Dr. Lara Finelli. European Polymer Journal. 2016;75:431-
460. doi:10.1016/j.eurpolymj.2016.01.016

8. Zhao C, He X, Zou G, Li J, Li J. Biodegradable poly(butylene succinate-co-butylene


dimerized fatty acid)s: Synthesis, crystallization, mechanical properties, and rheology.
Polymer Science - Series B. 2016;58(2):183-193. doi:10.1134/S1560090416020111

9. Wierckx N, Narancic T, Eberlein C, et al. Plastic Biodegradation: Challenges and


Opportunities. Consequences of Microbial Interactions with Hydrocarbons, Oils, and
Lipids: Biodegradation and Bioremediation. Published online 2019:333-361.
doi:10.1007/978-3-319-50433-9_23

10. Díaz A, Katsarava R, Puiggalí J. Synthesis, properties and applications of


biodegradable polymers derived from diols and dicarboxylic Acids: From Polyesters to
poly(ester amide)s. International Journal of Molecular Sciences. 2014;15(5):7064-
7123. doi:10.3390/ijms15057064

18
11. Kanemura C, Nakashima S, Hotta A. Mechanical properties and chemical structures of
biodegradable poly(butylene-succinate) for material reprocessing. Polymer
Degradation and Stability. 2012;97(6):972-980.
doi:10.1016/j.polymdegradstab.2012.03.015

12. BioPBSTM Technical Data Sheet – MC PP Germany GmbH, A Group of Mitsubishi


Chemical, 2020.

13. Sousa A, Patrício R, Terzopoulou Z, et al. Recommendations for replacing PET on


packaging, fiber, and film materials with biobased counterparts. Green Chemistry.
2021;23(22):8795-8820. doi:10.1039/d1gc02082j

14. Wcislek A, Olalla A, McClain A, et al. Enzymatic degradation of poly(butylene


succinate) copolyesters synthesized with the use of Candida antarctica lipase B.
Polymers (Basel). 2018;10(6). doi:10.3390/polym10060688

15. Stȩpień K, Miles C, McClain A, et al. Biocopolyesters of Poly(butylene succinate)


Containing Long-Chain Biobased Glycol Synthesized with Heterogeneous Titanium
Dioxide Catalyst. ACS Sustainable Chemistry and Engineering. 2019;7(12):10623-
10632. doi:10.1021/acssuschemeng.9b01191

16. Zhang Q, Song M, Xu Y, Wang W, Wang Z, Zhang L. Bio-based polyesters: Recent


progress and future prospects. Progress in Polymer Science. 2021;120:101430.
doi:10.1016/j.progpolymsci.2021.101430

17. Lubkowski K, Smorowska A, Grzmil B, Kozłowska A. Controlled-release fertilizer


prepared using a biodegradable aliphatic copolyester of poly(butylene succinate) and
dimerized fatty acid. Journal of Agricultural and Food Chemistry. 2015;63(10):2597-
2605. doi:10.1021/acs.jafc.5b00518

18. Jäger A, Gromadzki D, Jäger E, et al. Novel “soft” biodegradable nanoparticles


prepared from aliphatic based monomers as a potential drug delivery system. Soft
Matter. 2012;8(16):4343-4354. doi:10.1039/c2sm07247e

19. Xu J, Guo BH. Poly(butylene succinate) and its copolymers: Research, development
and industrialization. Biotechnology Journal. 2010;5(11):1149-1163.
doi:10.1002/biot.201000136

20. Jost V. Packaging related properties of commercially available biopolymers - An


overview of the status quo. Express Polymer Letters. 2018;12(5):429-435.
doi:10.3144/expresspolymlett.2018.36

21. Francisco A, Lorevice M, Claro P, Gouveia R. Comprehensive study of cellulose


nanocrystals acetylation effects on poly (butylene adipate-co-terephthalate)
nanocomposite films obtained by solvent casting and heat pressing. Industrial Crops
and Products. 2022;177:114459. doi:10.1016/j.indcrop.2021.114459

19
22. Li M, Jia Y, Shen X, et al. Investigation into lignin modified PBAT/thermoplastic
starch composites: Thermal, mechanical, rheological and water absorption properties.
Industrial Crops and Products. 2021;171:113916. doi:10.1016/j.indcrop.2021.113916

23. Wadaugsorn K, Panrong T, Wongphan P, Harnkarnsujarit N. Plasticized


hydroxypropyl cassava starch blended PBAT for improved clarity blown films:
Morphology and properties. Industrial Crops and Products. 2022;176:114311.
doi:10.1016/j.indcrop.2021.114311

24. Ferreira F, Cividanes L, Gouveia R, Lona L. An overview on properties and


applications of poly(butylene adipate-co-terephthalate)–PBAT based composites.
Polymer Engineering and Science. 2019;59(s2):E7-E15. doi:10.1002/pen.24770

25. Zaverl M, Valerio O, Misra M, Mohanty A. Study of the effect of processing conditions
on the co-injection of PBS/PBAT and PTT/PBT Blends for parts with increased bio-
content. Journal of Applied Polymer Science. 2015;132(2):1-11.
doi:10.1002/app.41278

26. Su S, Duhme M, Kopitzky R. Uncompatibilized PBAT/PLA blends: manufacturability,


miscibility and properties. Materials. 2020;13:4897. Doi:10.3390/ma13214897

20

You might also like