You are on page 1of 7

Polymer Degradation and Stability 211 (2023) 110335

Contents lists available at ScienceDirect

Polymer Degradation and Stability


journal homepage: www.journals.elsevier.com/polymer-degradation-and-stability

Enzymatic hydrolysis of poly(butylene adipate-co-terephthalate) by


Fusarium solani cutinase
Wen Lin a, Yujin Zhao a, Tingting Su a, *, Zhanyong Wang b, *
a
School of Petrochemical Engineering, Liaoning Petrochemical University, Fushun 113001, China
b
College of Bioscience and Biotechnology, Shenyang Agricultural University, Shenyang 110866, China

A B S T R A C T

The degradation behavior of poly(butylene adipate-co-terephthalate) (PBAT) films by Fusarium solani cutinase (FsC) applied at different concentrations was inves­
tigated. The degraded PBAT films were characterized by scanning electron microscopy, X-ray diffraction, attenuated total reflectance Fourier transform infrared
spectroscopy, and thermogravimetry. The results showed that FsC was able to degrade not only the amorphous region but also the crystalline region of PBAT films.
The PBAT degradation activity of FsC was concentration-dependent. After a reaction time of 120 h, the weight loss of PBAT films degraded by FsC of a high con­
centration reached 76%, and to a significant extent, the surface of the film was degraded and the crystallinity decreased. In contrast, when FsC was present in a low
concentration, its degradation effect on PBAT films was less apparent. The thermal stability of PBAT showed a decreasing trend with the extension of FsC hydrolysis
time. Although the production of monomers and oligomers during the hydrolysis of PBAT by different concentrations of FsC was significantly different, the products
were identical after 120 h of hydrolysis. In addition, FsC concentration and degradation time had no significant effect on the molecular weight of degraded PBAT.

1. Introduction to catalyze the hydrolysis of some synthetic polyesters [12,13]. It has an


open and shallow active site without a "lid"-structure that accommo­
The increasing commercialization within modern society has also dates the polyester chain molecule and affects the contact of the poly­
created an increasing dependency on plastic. Especially in the global ester with the active serine at the catalytic triad (Ser-His-Asp) [14].
fight against COVID-19 in recent years, disposable masks, gloves, and In our previous works, we verified that the cutinase purified from
protective clothing mainly made of plastic have played an important Fusarium solani (FsC) could biodegrade aliphatic polyesters containing
role in controlling the spread of the virus [1]. In the environment, adipic acid or 1,4-butanediol as monomers [15]. The objective of this
however, the used plastic waste cannot be completely degraded and will current work was to comprehensively analyze the effect of enzymatic
cause serious environmental pollution. The resulting microplastics pose PBAT hydrolysis by FsC on the degradation characteristics of PBAT. To
a serious threat to human health and marine ecosystems [2–4]. The this end, the changes in the surface morphology of PBAT film and the
promotion and application of bio-based materials and degradable plas­ chemical structure of PBAT were determined after enzymatic hydrolysis
tics can alleviate the above problems. by scanning electron microscopy (SEM), attenuated total reflectance
Poly(butylene adipate-co-terephthalate) (PBAT) is a semi-crystalline Fourier transform infrared (ATR-FTIR) spectroscopy, thermogravimetric
aliphatic-aromatic co-polyester, which is produced by polycondensation analysis (TGA), and powder X-ray diffraction (XRD). The products of
from adipic acid, terephthalic acid, and 1,4-butanediol[5,6]. The me­ PBAT degradation by FsC were also analyzed by mass spectrometry
chanical properties of PBAT are similar to low-density polyethylene (MS). This study reveals the difference of different concentrations of FsC
(LDPE) and can replace LDPE as plastic packaging in the food industry on the same substrate PBAT and lays the foundation for achieving
and plastic mulch in the agricultural field [7]. Although there are ester controlled degradation of PBAT.
bonds in PBAT that are theoretically easy to be hydrolyzed by microbial
enzymes, PBAT is more difficult to be degraded than poly­ 2. Materials and methods
hydroxyalkanoates, which are aliphatic polyesters [8]. According to
previous studies, cutinase can effectively hydrolyze PBAT [9,10]. Cuti­ 2.1. Materials
nase is an α/β-hydrolase that belongs to the extracellular serine esterase
family and is widespread in bacteria and fungi [11]. It has been reported PBAT (TH801T) with an average molecular weight of 18,000 was

* Corresponding authors.
E-mail addresses: sutingting1978@126.com (T. Su), wangzy125@gmail.com (Z. Wang).

https://doi.org/10.1016/j.polymdegradstab.2023.110335
Received 28 November 2022; Received in revised form 7 February 2023; Accepted 10 March 2023
Available online 12 March 2023
0141-3910/© 2023 Elsevier Ltd. All rights reserved.
W. Lin et al. Polymer Degradation and Stability 211 (2023) 110335

obtained from Xinjiang Blue Ridge Tunhe Sci. & Tech. Co., Ltd, China. 2.6. Gel permeation chromatography (GPC)
FsC was prepared from the zymotic fluid of recombinant Pichia pastoris
[16]. All chemicals were analytical-grade reagents and acquired from GPC was used for the analysis of the average molecular weights and
Sinopharm Chemical Reagent Co., Ltd. molecular weight distribution of synthesized polyesters. GPC was con­
ducted with a Waters 1515 Isocratic HPLC Pump (Milford, MA, USA). A
Waters 1515 refractive index detector was used with a temperature
2.2. Preparation of PBAT film
controller. The polyesters were dissolved in chloroform to prepare a 1.5
mg/mL solution. After being fully dissolved, the solution was filtered by
After drying at 80 ◦ C for 12 h, PBAT particles were fed into a single-
0.22 μm membrane, and then 10 lL filtrate was directly injected for
screw extruder operated with a coat hanger die at 180 ◦ C and 50 rpm
testing. Waters Styragel HT3 (7.8 × 300 mm) and HT4 column (7.8 ×
with rollers at 30 rpm and 55 ◦ C to generate flat films with an approx­
300 mm) were used in tandem, and different molecular masses of
imate thickness of 0.1 mm. The film was cut into 1 × 3 cm samples,
polystyrene were utilized as standard. The pore size of the column was
which were dried at 50 ◦ C to constant weight.
10 lm. Other specific test conditions were detailed as follows: column
temperature is 35 ◦ C, column pressure is 1600 psi, mobile phase used
2.3. Enzymatic hydrolysis was chloroform and flow velocity is 0.8 mL/min.

PBAT films were incubated in 10 mL potassium phosphate buffer (pH 3. Results and discussion
7.2) with two different cutinase concentrations (3.6 U/mL and 360 U/
mL) at 37◦ C. The films were taken out every 24 h and soaked with 2% 3.1. Weight loss of PBAT films
aqueous sodium dodecyl sulfate solution for 2 h, then thoroughly
washed with deionized water, and dried under reduced pressure to a The weight loss curves of PBAT films degraded by FsC of different
constant weight. The dry weights of films before and after hydrolysis concentrations are shown in Fig. 1. Whether under the action of low or
were determined to calculate the weight loss of PBAT films using Eq. (1). high FsC concentration, the weight loss of PBAT films both increased
Wbefore − Wafter with hydrolysis time, but the hydrolysis effects of FsC concentrations on
Wloss (%) = × 100% (1) the films were different. After degradation for 120 h, the weight loss of
Wbefore
PBAT films hydrolyzed by low-concentration FsC only reached 16.6%,
where Wbefore is the dry weight (g) before hydrolysis, and Wafter is the dry while that of PBAT films degraded by high-concentration FsC was as
weight (g) after hydrolysis. high as 76%. This indicates that the hydrolysis of PBAT by FsC is
dependent on enzyme concentration, which is similar to the dependence
of the degradation efficiency of poly(lactic acid) on the concentration of
2.4. Characterization of PBAT films after enzymatic hydrolysis proteinase K [17]. The observed influence of enzyme concentration can
also explain the slow degradation of PBAT in the natural environment
2.4.1. SEM because the concentration of enzymes produced by PBAT-degrading
The micromorphology of PBAT films was observed via SEM using a microorganisms in the natural environment is insufficient to achieve
Hitachi SU8010 scanning electron microscope (Japan) at 20 kV accel­ rapid PBAT hydrolysis. The control samples without enzyme treatment
eration. Before testing, the surface of PBAT films was sprayed with a thin did not show any significant weight loss (the data are not represented in
layer of gold to conduct electrons. Fig. 1). It also indicates that the weight loss in the experimental group
was brought about by enzyme catalysis rather than direct hydrolysis.
2.4.2. ATR-FTIR spectroscopy The surface morphologies of PBAT films after degradation are shown
An Agilent Cary 660 FTIR spectrometer with a slide-on ATR acces­ in Fig. 2. The film was milky white before degradation, and the surface
sory (USA) was used for FTIR analysis. The analysis was conducted at a was smooth and complete. After degradation by low-concentration FsC,
resolution of 2 cm− 1 over a frequency range of 4000–400 cm− 1 and the the color, overall shape, and size of the films did not change
number of scans was 32.

2.4.3. XRD
The crystallinity of PBAT was investigated with a Bruker S8 Tiger
Advance X-ray diffractometer (Germany). The XRD patterns were ob­
tained within the range of 5◦ − 60◦ with a step size of 0.02◦ The XRD
instrument was operated at 40 kV and 40 mA using Cu Kα radiation.

2.4.4. TGA
The thermal decomposition behavior of PBAT films was determined
with Q600 thermal analyzer (TA Instruments, USA). Approximately 10
mg of each sample was heated in nitrogen atmosphere from room
temperature to 500 ◦ C at a rate of 10 ◦ C⋅min− 1.

2.5. Analysis of enzymatic PBAT hydrolysates

After removing the degraded PBAT films, the buffer solutions were
subjected to centrifugal ultrafiltration with a molecular weight limit of
3000 Da, and the fluid below the membrane was assayed by MS (MALDI-
TOF-MS; 4700 Proteomics Analyzer; Applied Biosystems, USA). Elec­
trospray ionization was used, and positive and negative ionization
modes were scanned simultaneously. The capillary voltage was 3.5 kV,
the sprayer voltage was 20 kV, the RF lens voltage was 0.5 V, and the Fig. 1. Weight loss of PBAT film hydrolyzed by FsC (Values are mean ± stan­
source temperature was 120 ◦ C. dard deviation, n = 3).

2
W. Lin et al. Polymer Degradation and Stability 211 (2023) 110335

Fig. 2. Surface morphology of PBAT films hydrolyzed by FsC.

significantly, but FsC eroded the surface of the PBAT film, resulting in a and the intensity of diffraction peaks showed a significant downward
thin film with a rough and dull surface. After the film was degraded by trend with increasing degradation time.
high-concentration FsC, obvious gaps and cracks appeared in the film.
With increasing degradation time, the film gradually became thinner, 3.4. ATR-FTIR spectroscopy
and the surface of the film became rougher. When degraded for 120 h,
the film surfaces were significantly damaged, and the shape was The changes caused by FsC on PBAT films were analyzed by ATR-
irregular. FTIR spectroscopy (Fig. 5). The absorption peak in the range of
3000–2800 cm− 1 represents the asymmetric stretching vibration of
3.2. Micro-morphological changes of PBAT films methylene (-CH2-) groups, the strong absorption peak at 1710 cm− 1 is
generated by the stretching vibration of ester carbonyl (-C=O) groups,
As shown in Fig. 3, SEM of the PBAT film before hydrolysis showed the absorption peak in the range of 1300–1100 cm− 1 corresponds to the
that its surface was smooth and flat. The damage caused to the PBAT stretching vibration of the -C-O- group of the terephthalic acid ester
films after being degraded by FsC of different concentrations was group in PBAT, and the absorption peak near 720 cm− 1 is the external
significantly different. After being incubated for 24 h with low- bending vibration of the aromatic C–H group of the terephthalate
concentration FsC, the surface of the film was slightly changed and moieties in PBAT [5]. The above absorption peaks are characteristic
became rough. The surface of the PBAT film degraded for 72 h presented peaks of PBAT.
small holes and increased roughness. After hydrolysis for 96 h, the After the PBAT film was degraded by low-concentration FsC, a new
surface showed cracks, and the degree of degradation was further absorption peak appeared near 3300 cm− 1 while shapes and positions of
increased. After being hydrolyzed for 120 h, many crevices appeared on other absorption peaks did not change significantly. As a result of main
the film, and their depth was increased. In contrast, the films hydrolyzed chain scission by hydrolysis at ester linkages, terminal alcohol and
by high-concentration FsC were severely eroded after 24 h, and uneven carboxylic acid groups are produced, so as hydrolysis progresses, an
bulges and multiple depressions appeared on the surface of the films. increase in OH groups should be observed in the FTIR spectra [20]. This
The cracks on the film hydrolyzed for 72–120 h became longer and is confirmed by the presence of the peak at 3300 cm− 1, identified as the
deeper, the number of holes and cracks increased, and the film surface hydroxyl vibration absorption peak which also appeared after PBAT was
was rougher. FsC can effectively destroy the structure of PBAT film, and degraded by high-concentration FsC. Moreover, the change of absorp­
the higher the concentration, the more significant the hydrolysis effect. tion peak near 1710cm− 1also reflects the changes of carboxylic acid
groups within the biodegradation.
3.3. XRD
3.5. TGA
The XRD patterns (Fig. 4a and b) of the PBAT films degraded by FsC
can reflect the changes in the crystalline properties of the PBAT films TGA was employed to monitor the degradation process in a tem­
(Fig. 4c). The diffraction peaks at 16.1, 17.5, 20.3, and 22.8◦ are char­ perature range of 50–600 ◦ C, and the results are displayed in Fig. 6. The
acteristic peaks of PBAT, corresponding to the (011), (010), (111), and PBAT films degraded by FsC had the same thermal stability and
(100) crystal planes, respectively [18]. Regardless of the FsC concen­ exhibited a one-step decomposition pattern. The thermal decomposition
tration used, the diffraction peak positions did not change, indicating temperatures of PBAT after degradation are listed in Table 1.
that the degradation of PBAT by FsC will not cause a change in the The thermal stability of PBAT after degradation changed with
crystal structure of PBAT. However, the intensity of diffraction peaks increasing hydrolysis time, and the thermal decomposition temperature
changed significantly after degradation, and the degree of change varied of PBAT films showed a decreasing trend due to the generation of end
for different crystal planes, which was consistent with the conclusion of groups that facilitated the thermal decomposition reaction of PBAT
Jia et al. that PBAT was degraded by Stenotrophomonas sp. YCJ1 [19]. [21]. A phase of relatively higher thermal decomposition temperature
When PBAT was degraded by low-concentration FsC for 24 h, the was observed during the degradation of low concentration FsC. This may
shape of diffraction peaks became wider and the peak areas decreased be due to the increase in crystallinity caused by the faster degradation of
compared with the control without enzyme treatment, indicating that the amorphous region of PBAT film at this stage. The TGA curve of PBAT
the crystallinity of PBAT was reduced. After degraded for 48–72 h, the degraded by high-concentration FsC showed a significant decrease in its
intensity of the PBAT diffraction peaks at 17.5 and 20.3◦ increased. This thermal decomposition temperature compared with low-concentration
may be due to the ordered arrangement of molecular chains in the FsC, which indicated that at the higher concentration, FsC had a stron­
crystalline region of PBAT, which is difficult to be eroded by FsC. This ger destructive force on PBAT and effected a faster polyester chain
would cause the amorphous regions to degrade faster than the crystal­ breakage.
line regions, thus leading to an increase in crystallinity. After being
degraded for 72–120 h, the regular molecular chains of the crystalline 3.6. Determination of PBAT hydrolysis products
region of PBAT were disrupted due to the long-term action of FsC, the
crystalline regions of PBAT were transformed into amorphous regions, After the incubation of PBAT film with FsC, hydrolysates released
and the crystallinity was further reduced. When high-concentration FsC into the supernatant were identified by MS. As shown in Table 2, the
acted on PBAT, the high enzyme concentration led to faster degradation, hydrolytic monomers B (1,4-butanediol) and T (terephthalic acid) were

3
W. Lin et al. Polymer Degradation and Stability 211 (2023) 110335

Fig. 3. SEM of PBAT films hydrolyzed by FsC.

4
W. Lin et al. Polymer Degradation and Stability 211 (2023) 110335

Fig. 4. XRD and crystallinity changes of PBAT films hydrolyzed by FsC (a, hydrolyzed by low concentration FsC; b, hydrolyzed by high concentration FsC; c,
crystallinity of PBAT films, Xc).

Fig. 5. ATR-FTIR of PBAT films hydrolyzed by FsC (a, hydrolyzed by low concentration FsC; b, hydrolyzed by high concentration FsC).

Fig. 6. TGA of PBAT films hydrolyzed by FsC (a, hydrolyzed by low concentration FsC; b, hydrolyzed by high concentration FsC).

detected in all tested samples, while monomer A (adipic acid) was not while BA was present in the prehydrolysis stage of low-concentration
detected in the early stage of hydrolysis. Monomer A was observed after FsC and disappeared after hydrolysis of 72 h Among the potential hy­
72 h of hydrolysis by high-concentration FsC and after 96 h of hydrolysis drolytic trimers BAB, ABA, TBT, BTB, and ABT, only BAB was not
by low-concentration FsC, which may imply that its production origi­ detected, which constitutes the same trimer product spectrum observed
nates from the enzymatic action on the oligomer rather than from the after fungal degradation [22] or enzymatic hydrolysis [23] of PBAT. It is
direct hydrolysis of PBAT. In addition, only dimer BT without any BA worth noting that no tetramers were found in all tested samples, while
was observed in the hydrolysis products of high-concentration FsC, low levels of pentamers were also detected. The hydrolysis products of

5
W. Lin et al. Polymer Degradation and Stability 211 (2023) 110335

Table 1 Table 3
Thermogravimetric characteristic of PBAT films hydrolyzed by FsC. Molecular weight of PBAT films hydrolyzed by FsC.
Degradation Time Low FsC High FsC Degradation Time Low FsC High FsC

T5% ( C)◦
T50% ( C)

T5% ( C)◦
T50% ( C)

Mw Mn Mw Mn

CK 378.7 418.7 378.7 418.7 CK 142,900 33,800 142,900 33,800


24 h 377.1 417.1 368.6 416.1 24 h 159,500 28,000 147,800 24,900
48 h 368.6 416.1 359.5 414.5 48 h 141,800 27,900 144,500 28,000
72 h 374.5 419.5 352.5 415.0 72 h 149,200 34,700 149,500 39,000
96 h 369.9 417.5 346.6 416.6 96 h 155,900 29,000 144,500 29,700
120 h 367.6 415.1 337.6 415.1 120 h 149,600 30,900 147,000 30,500

Note: T5% is decomposition temperature of 5%; T50% is decomposition temper­


ature of 50%. poly(butylene succinate) and Poly(hexylene succinate) by FsC, it was
also found that the molecular weight of the degraded polyesters did not
change significantly [27]. The GPC results support that the most likely
Table 2
Hydrolyzed monomers or oligomers of PBAT degraded by FsC.
mechanism of degradation of PBAT by FsC is surface erosion, since the
molecular weight of the PBAT film did not change during the experi­
Degradation time Monomers or oligomers from hydrolysis
ment. The surface erosion mechanism usually results in a reduction in
(h)
the size of the sample by forming pores or eroding edges on the surface
24 Low FsC B, T, BA, BT, BTB, TBT, BABTB/BTBAB, BABAB, of the polyester films [28]. This is also verified in Figs. 2 and 3.
BTBTB
High B, T, BT, BTB, TBT, BABTB/BTBAB, BABAB,
FsC BTBTB 4. Conclusions
48 Low FsC B, T, BA, BT, ABA, TBT, BABTB/BTBAB, BABAB,
BTBTB In this work, we investigated the changes of PBAT films after
High B, T, BT, BTB, TBT, BABTB/BTBAB, BABAB,
FsC BTBTB, BTBABT
degradation by FsC. The weight of PBAT films was significantly reduced
72 Low FsC B, T, BA, BT, BTB, TBT, BABTB/BTBAB, BABAB, after enzymatic degradation, and the surface and internal structures of
BTBTB PBAT films were destroyed. FsC effectively degraded the amorphous and
High B, A, T, BT, BTB, BABAB crystalline regions of PBAT, and the crystalline regions were trans­
FsC
formed into amorphous regions with the increase in degradation time.
96 Low FsC B, A, T, BT, BTB, TBT, BABAB, BTBTB
High B, A, T, BT, BTB, BABAB, The disruption of the PBAT structure led to a decrease in the thermal
FsC stability of partly degraded PBAT. The effect of a high concentration of
120 Low FsC B, A, T, BT, BTB, BABAB, FsC on PBAT was more significant compared with its employment at a
High B, A, T, BT, BTB, BABAB low concentration, indicating that the degradation of PBAT was
FsC
dependent on enzyme concentration. The production of monomers and
Note: B, 1,4-butanediol; A, adipic acid; T, terephthalic acid. oligomers during the hydrolysis of PBAT by different concentrations of
FsC was significantly different, but the types of product were identical
PBAT obtained with FsC are mainly B, T, BT, and ABT. Marten et al. after 120 h of hydrolysis. It is noteworthy that the production of
reported that some lipases and Thermobifida fusca hydrolase (TfH) are monomer A significantly lags behind that of monomers B and T. The
less effective in cleaving the ester bond in PBAT close to T [24]. The molecular weight of the degraded PBAT films did not change during the
PBAT hydrolase PBATHRf from Rhodococccus fascians preferentially experiment, which proved that the degradation mechanism of FsC on
cleaves the ester bond between B and A rather than the ester bond be­ PBAT films was realized by surface erosion.
tween B and T[23]. And the PBAT hydrolase PBATHBp derived from
Bacillus pumilus hydrolyses ester bonds between B and T at much slower CRediT authorship contribution statement
rates than ester bonds between adipate and butanediol[10]. The hy­
drolysis of PBAT by FsC behaves differently, which may be related to the Wen Lin: Conceptualization, Methodology, Investigation, Formal
type of enzyme. The hydrolysis of PBAT molecules occurs mainly at the analysis, Writing – original draft, Writing – review & editing. Yujin
ester bond between A and T. Zhao: Conceptualization, Methodology, Investigation, Formal analysis.
It was previously reported that the hydrolysis of PBAT molecules Tingting Su: Supervision, Formal analysis, Writing – review & editing.
mainly takes places on the ester bond located between the terephthalate Zhanyong Wang: Methodology, Supervision, Project administration,
and the adipate groups [25]. Deshoulles et al. in their study on the aging Funding acquisition.
hydrolysis of PBAT in water at 80–100 ◦ C also found that hydrolysis also
takes place on the ester located between the terephthalate and adipate
Declaration of Competing Interest
groups [26]. And this phenomenon was verified in our enzymatic hy­
drolysis of PBAT using FsC. Oligomers containing ester bonds within the
The authors declare that they have no known competing financial
terephthalate and adipate groups were not present in the degradation
interests or personal relationships that could have appeared to influence
products, indicating that such ester bonds preferentially underwent
the work reported in this paper.
hydrolysis.
Data availability
3.7. GPC
No data was used for the research described in the article.
The weight-average molecular weight (Mw) and number-average
molecular weight (Mn) and of PBAT films after degraded by FsC are
listed in Table 3. No significant changes in the molecular weight of PBAT Acknowledgments
were observed with increasing degradation time, which implies that the
degradation process occurs uniformly on the film surface. In our previ­ This work was supported by National Natural Science Foundation of
ous studies on the enzymatic degradation of poly(ethylene succinate), China (Grant No. 32270117) and the Talent Program of Shenyang

6
W. Lin et al. Polymer Degradation and Stability 211 (2023) 110335

Agricultural University (Grant No. 2021Y001). [14] S. Joo, I.J. Cho, H. Seo, H.F. Son, H.Y. Sagong, T.J. Shin, S.Y. Choi, S.Y. Lee, K.
J. Kim, Structural insight into molecular mechanism of poly(ethylene
terephthalate) degradation, Nat. Commun. 9 (2018) 382.
References [15] Z. Bai, K. Shi, T. Su, Z. Wang, Correlation between the chemical structure and
enzymatic hydrolysis of Poly(butylene succinate), Poly(butylene adipate), and Poly
[1] J.I. Kwak, Y.J. An, Post COVID-19 pandemic: biofragmentation and soil (butylene suberate), Polym. Degrad. Stab. 158 (2018) 111–118.
ecotoxicological effects of microplastics derived from face masks, J. Hazard. Mater. [16] X. Hu, Z. Gao, Z. Wang, T. Su, L. Yang, P. Li, Enzymatic degradation of poly
416 (2021), 126169. (butylene succinate) by cutinase cloned from Fusarium solani, Polym. Degrad. Stab.
[2] S. Maity, S. Banerjee, C. Biswas, R. Guchhait, A. Chatterjee, K. Pramanick, 134 (2016) 211–219.
Functional interplay between plastic polymers and microbes: a comprehensive [17] K. Yamashita, Y. Kikkawa, K. Kurokawa, Y. Doi, Enzymatic degradation of poly(L-
review, Biodegradation 32 (2021) 487–510. lactide) film by proteinase K: quartz crystal microbalance and atomic force
[3] G.G.N. Thushari, J.D.M. Senevirathna, Plastic pollution in the marine environment, microscopy study, Biomacromolecules 6 (2005) 850–857.
Heliyon 6 (2020) e04709. [18] Z. Gan, K. Kuwabara, M. Yamamoto, H. Abe, Y. Doi, Solid-state structures and
[4] H.A. Leslie, M.J.M. Van Velzen, S.H. Brandsma, A.D. Vethaak, J.J. Garcia-Vallejo, thermal properties of aliphatic–aromatic poly(butylene adipate-co-butylene
M.H. Lamoree, Discovery and quantification of plastic particle pollution in human terephthalate) copolyesters, Polym. Degrad. Stab. 83 (2004) 289–300.
blood, Environ. Int. 163 (2022), 107199. [19] H. Jia, M. Zhang, Y. Weng, Y. Zhao, C. Li, A. Kanwal, Degradation of poly(butylene
[5] A. Kanwal, M. Zhang, F. Sharaf, C. Li, Enzymatic degradation of poly (butylene adipate-co-terephthalate) by Stenotrophomonas sp. YCJ1 isolated from farmland
adipate co-terephthalate) (PBAT) copolymer using lipase B from Candida antarctica soil, J. Environ. Sci. 103 (2021) 50–58.
(CALB) and effect of PBAT on plant growth, Polym. Bull. 79 (2022) 9059–9073. [20] T. Kijchavengkul, R. Auras, M. Rubino, S. Selke, M. Ngouajio, R.T. Fernandez,
[6] Q. Deshoulles, M.L. Gall, S. Benali, J.M. Raquez, C. Dreanno, M. Arhant, D. Priour, Biodegradation and hydrolysis rate of aliphatic aromatic polyester, Polym. Degrad.
S. Cerantola, G. Stoclet, P.Y.L. Gac, Hydrolytic degradation of biodegradable poly Stab. 95 (2010) 2641–2647.
(butylene adipate-co-terephthalate) (PBAT)-Towards an understanding of [21] D.N. Bikiaris, G.P. Karayannidis, Calorimetric study of diepoxide chain-extended
microplastics fragmentation, Polym. Degrad. Stab. 205 (2022), 110122. poly(ethylene terephthalate), J. Therm. Anal. Calorim. 54 (3) (1998) 721–729.
[7] P. Soulenthone, Y. Tachibana, F. Muroi, M. Suzuki, N. Ishii, Y. Ohta, K. Kasuya, [22] K. Kasuya, N. Ishii, Y. Inoue, K. Yazawa, T. Tagaya, T. Yotsumoto, J. Kazahaya,
Characterization of a mesophilic actinobacteria that degrades poly(butylene D. Nagai, Characterization of a mesophilic aliphatic–aromatic copolyester-
adipate-co-terephthalate), Polym. Degrad. Stab. 181 (2020), 109335. degrading fungus, Polym. Degrad. Stab. 94 (8) (2009) 1190–1196.
[8] H. Sashiwa, R. Fukuda, T. Okura, S. Sato, A. Nakayama, Microbial degradation [23] P. Soulenthone, Y. Tachibana, M. Suzuki, T. Mizuno, Y. Ohta, K.-i. Kasuya,
behavior in seawater of polyester blends containing poly(3-hydroxybutyrate-co-3- Characterization of a poly(butylene adipate-co-terephthalate) hydrolase from the
hydroxyhexanoate) (PHBHHx), Mar Drugs 16 (2018) 34. mesophilic actinobacteria Rhodococcus fascians, Polym. Degrad. Stab. 184 (2021),
[9] V. Perz, K. Bleymaier, C. Sinkel, U. Kueper, M. Bonnekessel, D. Ribitsch, G. 109481.
M. Guebitz, Substrate specificities of cutinases on aliphatic–aromatic polyesters [24] E. Marten, R.J. Müller, W.D. Deckwer, Studies on the enzymatic hydrolysis of
and on their model substrates, N Biotechnol 33 (2016) 295–304. polyesters. II. Aliphatic–aromatic copolyesters, Polym. Degrad. Stab. 88 (3) (2005)
[10] F. Muroi, Y. Tachibana, P. Soulenthone, K. Yamamoto, T. Mizuno, T. Sakurai, 371–381.
Y. Kobayashi, K. Kasuya, Characterization of a poly(butylene adipate-co- [25] F.V. Ferreira, L.S. Cividanes, R.F. Gouveia, L.M.F. Lona, An overview on properties
terephthalate) hydrolase from the aerobic mesophilic bacterium Bacillus pumilus, and applications of poly(butylene adipate-co-terephthalate)–PBAT based
Polym. Degrad. Stab. 137 (2017) 11–22. composites, Polym. Eng. Sci. 59 (s2) (2019) E7–E15.
[11] A. Su, A. Shirke, J. Baik, Y. Zou, R. Gross, Immobilized cutinases: preparation, [26] Q. Deshoulles, M.L. Gall, S. Benali, J.M. Raquez, C. Dreanno, M. Arhant, D. Priour,
solvent tolerance and thermal stability, Enzyme Microb. Technol. 116 (2018) S. Cerantola, G. Stoclet, P.Y.L. Gac, Hydrolytic degradation of biodegradable poly
33–40. (butylene adipate-co-terephthalate) (PBAT) - Towards an understanding of
[12] L.F. Ping, X.Y. Chen, X.L. Yuan, M. Zhang, Y.J. Chai, S.D. Shan, Application and microplastics fragmentation, Polym. Degrad. Stab. 205 (2022), 110122.
comparison in biosynthesis and biodegradation by Fusarium solani and Aspergillus [27] Z. Bai, Y. Liu, T. Su, Z. Wang, Effect of hydroxyl monomers on the enzymatic
fumigatus cutinases, Int. J. Biol. Macromol. 104 (2017) 1238–1245. degradation of Poly(ethylene succinate), Poly(butylene succinate), and Poly
[13] S. Sulaiman, S. Yamato, E. Kanaya, J.J. Kim, Y. Koga, K. Takano, S. Kanaya, (hexylene succinate), Polymers 10 (1) (2018) 90. Basel.
Isolation of a novel cutinase homolog with polyethylene terephthalate-degrading [28] K. Shi, T. Su, Z. Wang, Comparison of poly(butylene succinate) biodegradation by
activity from leaf-branch compost by using a metagenomic approach, Appl. Fusarium solani cutinase and Candida antarctica lipase, Polym. Degrad. Stab. 164
Environ. Microbiol. 78 (5) (2012) 1556–1562. (2019) 55–60.

You might also like