Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Polymers and the Environment

https://doi.org/10.1007/s10924-023-02979-8

ORIGINAL PAPER

All Green Microwave Assisted 99% Depolymerisation of Polyethylene


Terephthalate into Value Added Products via Glycerol Pre‑treatment
and Hydrolysis Reaction
Muhammad Azeem1 · Olivia A. Attallah1,2 · Cuneyt Erdinc Tas1 · Margaret Brennan Fournet1

Accepted: 21 June 2023


© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2023

Abstract
Energy-efficient and fast depolymerisation technologies present as new sustainable and green recycling routes for achieving
a circular economy for plastics. Herein, we present a highly efficient 2-step microwave-based (MW) degradation of polyeth-
ylene terephthalate (PET). Initially, a MW-assisted pre-treatment was evaluated using glycerol as a non-toxic reagent for the
conversion of PET into a modified form that makes it easily depolymerised. Box Behnken Design was employed to determine
the optimised pre-treatment conditions attaining maximum PET weight loss and favourable crystallinity and carbonyl indices
for the pre-treated PET. Glycerol of 12 mL volume and 3 min of 182W MW irradiation resulted in 11% PET weight loss at
onset temperature of degradation and gave rise to carbonyl index up to 4.22 and 33% crystallinity of pre-treated PET. MW
assisted hydrolysis of the pre-treated PET was then performed in the presence of sodium bicarbonate and ethylene glycol as
depolymerizing agents. Within 3 min, the proposed depolymerisation methodology provided 99.9% conversion of PET into
79.1% terephthalic acid (TPA), 17.6% monohydroxyethyl terephthalate (MHET), and 1.8% bis (2-hydroxyethyl) terephtha-
late (BHET). The obtained TPA was separated from the monomers mixtures and its purification was evaluated via different
characterization techniques against a standard TPA. A purity of 95%, 82.4 APHA colour value, 645.3 mgKOH/g acid number
and acceptable heavy metal content indicated that the purified TPA can be repolymerised as virgin PET.
Graphical Abstract

Keywords Depolymerisation · Polyethylene terephthalate · Microwave · Recycling · Monomers · Glycerol pre-treatment ·


Hydrolysis

Extended author information available on the last page of the article

13
Vol.:(0123456789)
Journal of Polymers and the Environment

Introduction scale require further improvements due to their huge


energy inputs, and slow reaction rates [13].
Polyethylene terephthalate (PET) belongs to the class of Microwave (MW) technology has shown great potential
engineering polymers which have a broad range of appli- for addressing the problems of energy loss and long reaction
cations because of their extensive corrosion resistance runs associated with the depolymerisation of PET. Several
and thermal stability [1]. According to recent studies, studies have been conducted to evaluate PET depolymerisa-
the production of PET was doubled resulting in a growth tion efficiency via MW-assisted glycolysis and hydrolysis
rate of more than 5% per annum and the elevation of reactions in combination or separately [14]. In one study,
plastic waste up to 8 million tons every year [2–5]. In the PET glycolysis reaction time was decreased by 21 h from
recent years, there has been intensive care in the devel- the original time (24 h) to 3 h in the presence of an oxalate-
opment of scalable and sustainable recycling methodolo- bridged binuclear iron (III) ionic liquid under MW irradia-
gies to reduce plastic accumulation in nature to close the tion. The overall yield of monomer (BHET) was 80% [15]. In
loop for a circular economy. Chemical recycling as one another study, MW-assisted hydrolysis was conducted using
of these methodologies has shown great potential to pro- deep eutectic solvent (DES) based on m-cresol and choline
duce polymers with higher quality and economic aspects. chloride (ChCl) in combination with different compositions
The importance of chemical recycling lies in its ability to of NaOH. The maximum yield of TPA monomer (91.55%)
convert polymers into their basic building blocks that can was achieved within 92 s of MW irradiation together with
be repolymerised into virgin forms or upcycled into other DES and 10% NaOH [16]. In our previous work we also
useful products. There are two subcategories for chemical studied the sequential glycolysis-hydrolysis depolymeri-
recycling which are distinguished based on the reaction sation of PET using ChCl-urea and ChCl-thiourea-based
systems involved such as transesterification-based and DESs and evaluated the conversion efficiency of PET into
esterification-hydrolysis-based [6, 7]. In general, commer- value-added monomers. Overall, 99% PET conversion into
cial PET production from TPA as the monomer follows the monomers was achieved within 3 min with a yield of TPA
production line, which consists of the extraction of raw (80.66), MHET (34.79), and BHET (0.59) [10].
materials (p-xylene) from petroleum/gas, and oxidation Nevertheless, MW assisted depolymerisation of PET as a
of p-xylene via a proper catalyst system. The oxidation promising and suitable for large-scale implementation still
reaction of p-xylene produces by-products such as p-toluic need constant developments to replace the caustic depoly-
acid and p-formyl benzoic acid, resulting in the appear- merising agents which cause corrosion within reaction ves-
ance of impurities in the TPA products. To overcome this sels with safer ones. Therefore, it is imperative to develop
problem, the manufacturing line also includes additional MW assisted depolymerisation treatments containing green
crystallisation/separation steps. Moreover, this production depolymerising agents to improve the sustainability of PET
step involves harsh reaction conditions such as high pres- recycling and reduce carbon footprints associated with con-
sure and temperature [8]. In contrast, chemical recycling ventional depolymerisation methodologies. For instance,
of post-consumer PET provides a straight forward route Xie et al. [17] used zinc acetate to depolymerise PET into
to produce monomers (such as TPA) by allowing them, BHET. The resultant yield of the monomer was 85% under
to be converted directly from waste plastic as a source the reaction conditions of ­196◦C temperature, 3 h time, 1%
of raw material instead of oil. Thus, many commercial weight ratio of catalyst to PET, and 5(w/w) weight ratio of
TPA production steps can be eliminated by instant and EG to PET. In another study, sodium sulfate was used as a
direct production through postconsumer PET. Therefore catalyst to depolymerise PET into BHET under an excess
chemical recycling of postconsumer PET significantly amount of EG. At 1­ 96◦C temperature and reaction time of
contributes to conserving resources, reducing waste, and 8 h, 65% selectivity for BHET was achieved [18]. However,
developing an overall sustainable environment [9]. Among most of these conventional methodologies are performed
the different conventional chemical recycling techniques, under harsh conditions. Therefore, it is necessary to develop
glycolysis and hydrolysis appear feasible for large-scale greener methodologies that operate under mild conditions
applications [10]. In the glycolysis of PET, bis (2-hydrox- to improve the scalability and sustainability of recycling.
yethyl) terephthalate (BHET) monomer is produced as One of these suggested green reagents is sodium bicar-
a result of replacing ester linkages of PET by hydroxyl bonate ­(NaHCO3), a monosodium salt of carbonic acid
groups [11]. In a typical hydrolysis processing of PET, with alkalising characteristics. It is white powder and crys-
terephthalic acid (TPA) and ethylene glycol (EG) are gen- talline, commonly known as baking soda, used to relieve
erated as depolymerised products in the presence of acid/ acid indigestion [19]. Recently, ­NaHCO3 has shown great
base or steam [12]. Nevertheless, the implementation and potential for the glycolysis and hydrolysis recycling of PET.
practice of these chemical methods on a large industrial For instance, Dorin et al. [20] used ­NaHCO3 as a catalyst in
the depolymerisation of waste PET via hydrolysis reaction.

13
Journal of Polymers and the Environment

Within 150 min in a batch reactor, a 95% TPA yield was depolymerise PET into its value-added monomers. The TPA
achieved at ­196◦C temperature and 3 MPa reactor pressure obtained from the depolymerisation process was then sepa-
conditions. In another study, Sun cong-hao et al.[21] imple- rated, purified and validated for repolymerisation into PET
mented alcohol-alkali hydrolysis for the preparation of TPA in terms of organic structure, metallic content, colour, and
from PET waste bottles in the presence of ­NaHCO3. 94–97% acid number.
TPA was obtained with an overall 98% PET conversion
under reaction times (60–75 min), temperature (170–180◦C),
mass ratio of N ­ aHCO3 to PET of 1:1.1; and different volume Experimental
ratios of water to PET mass/EG to PET mass.
Glycerol on the other hand is also an inexpensive, green, Materials
biobased, organic molecule with a huge number of appli-
cation in chemistry and PET depolymerisation reactions PET polymer pellets were purchased from Alpek Polyes-
[22, 23]. For example, Dzhabarov et al. [23] used biodiesel ter UK Limited and micronised into fine powder by using
wastes containing glycerol and potassium salts to study PET a centrifugal- ZM200 miller. Analytical grades of sodium
depolymerisation in a glass reactor. Complete conversion of bicarbonate (98%), glycerol (99%), ethylene glycol (99%),
PET was achieved at 190–195°C temperature within 90 min. N,N-dimethyl acetamide (HPLC, ≥ 99.9%), ethyl alcohol
In another study, glycerol-based DES also delivered excep- (anhydrous, ≥ 99.5%) were obtained from Sigma Aldrich
tionally high PET conversion at short reaction times. The (UK).
PET conversion achieved was 38–61% at 1.3–1.6 min [24].
This could be associated with the efficient H-bond action MW Assisted Pre‑treatment of PET
between glycerol and urea that preferentially attacked the
ester group in PET. Furthermore efficient MW absorption Experiments were performed by adding 1 g of PET pow-
properties and increasing thermal heating response of glyc- der in different volumes ratios of glycerol. After stirring
erol under MW irradiations could also result in improved for 5 min, suspensions were kept under MW irradiation in
catalytic performance in the different depolymerisation reac- a 1200 W domestic MW oven at specific MW irradiation
tions [25, 26]. powers and times. After MW exposure, the pre-treated PET
Noting the effect of glycerol as efficient green MW was separated via filtration, followed by washing with dis-
absorbers and ­NaHCO3 in combination with EG as green tilled water prior to overnight oven drying at 65 °C. The MW
depolymerising agents. EG is used in the nonaqueous assisted pre-treatment of PET was optimised using a three-
hydrolysis of PET as a reaction solvent to enhance the reac- level, three factor BBD (Design Expert,11.1.0.1, Stat-Ease
tion rates. It provides hydroxyl groups which react with ester Inc., Minneapolis, MN). Three factors; MW irradiation time
bonds, causing cleavage and the resulting depolymerisation ­(X1) (min), MW power (­ X2) (W), and glycerol volume (­ X3)
of PET [27]. For example, Hu et al. [28] studied the depoly- (mL) were studied by running 15 experiments in triplicates
merisation of PET using NaOH in the presence and absence and mean data was reported. The responses were obtained as
of EG. In the absence of EG, 96% depolymerisation was crystallinity index calculated from DSC data, weight loss %
achieved at 6­ 0°C with a reaction time of 7 h using NaOH at onset temperature (­ T0) of degradation from TGA results,
in methanol. However, with the addition of EG(5 mL) to and carbonyl index from FTIR spectra (Table 1).
NaOH, the complete depolymerisation of PET into TPA was
achieved within 15 min. In a similar study [29], ethereal MW‑Assisted Hydrolysis of PET Obtained After
solvents (tetrahydrofuran, dioxane, dimethoxyethane) were Glycerol Pre‑treatment
combined with EG and NaOH solutions to accelerate PET
depolymerisation. All the PET pellets were converted to dis- Hydrolysis reaction of PET was carried out on both
odium terephthalate within 15 min at 1­ 80◦C. This indicates untreated and pre-treated PET to determine the necessity of
that EG and ­NaHCO3 can serve as good alternatives to con- the MW assisted pre-treatment step. Following the procedure
ventional corrosive and harsh reagents which require longer
times for PET depolymerisation reactions. In this research
Table 1  BBD levels and variables for PET pre-treatment process
study, fast and all-green sequential PET depolymerisation
technique comprising MW assisted pre-treatment followed Independent variables Level Constrains
by hydrolysis reaction was employed. The MW assisted pre- −1 0 1
treatment was conducted in the presence of glycerol, and
X1: MW time (min) 2 3 4 In the range
the process was optimised using Box-Behnken design. In
X2: MW power (W) 160 330 500 In the range
the hydrolysis reaction, the pre-treated PET was exposed to
X3: Volume of glycerol (mL) 8 12 16 In the range
MW treatment while using N ­ aHCO3 in EG to completely

13
Journal of Polymers and the Environment

developed in our previous work [10], untreated PET and pre- The thermal characteristics of PET samples were attained
treated PET were mixed separately in 10%(w/v) ­NaHCO3 by a differential scanning calorimetry (DSC)(Perkin Elmer
in 20 mL EG and stirred for 5 min. The prepared suspen- 4000) under nitrogen flow. Each sample weighing between
sions were then exposed to MW irradiations for 3 min in a 5–6 mg were kept in aluminium pan and heated from 30 to
MW oven at 350 W power. After MW treatment, dissolved ­275°C at a rate of 10 °C per min. The crystallinity index was
contents of PET were precipitated by adding distilled water calculated as follows [34];
in the suspensions. Upon filtration, the filtrate consisting of
(4)
( )
soluble monomers was characterised by high performance Crystallinity index = ΔHm ∕WΔHm0 × 100
liquid chromatography (HPLC). Pre-treated PET from both
where heat of fusion of PET samples is ∆Hm ­(jg−1), weight
samples was dried at 6­ 5◦C overnight. The precipitation of
fraction is W(g), and heat of fusion of completely crystalline
TPA from the filtrate was carried out using 2 mL hydro-
PET (140 ­jg−1) [16]is denoted by ∆Hm0.
chloric acid (34%) followed by washing with distilled water
The weight loss % of PET was analysed by using ther-
and drying at 6­ 5◦C overnight. Depolymerisation of PET was
mogravimetric analyser (TGA). All the measurements were
determined using the following equation [10]:
taken at onset temperature of degradation ­(To). The MW
PET conversion (%) = (1 − (Residual PET weight) treated samples were kept in a standard pan and heated
(1) (30–600°C) at the rate of 1­ 0◦C per min. The nitrogen flow of
∕(Initial PET weight) × 100)
50 mL per min was maintained for all the experiments. The
The selectivity measurement for the produced BHET, following equation was used to calculate PET weight loss %;
MHET, and TPA was extracted from HPLC chromatograms Weight loss (%) = (100 − wieght% of PET atTo) (5)
by quantifying peak area normalisation method and the fol-
lowing equation was used to calculate the yield of TPA [30]: Identification and quantification of soluble depolymerised
products from PET was determined using HPLC (Alliance
TPA yield (%) = ((PET conversion(%) waters, UK) with a flow rate of 1.2 mL/min in a C-18 col-
(2)
×TPA selectivity(%)))∕100 umn (4.6*250, 5 μm, Nucleodur, Germany). The injection
volume for each specimen was 10ul at ­30°C and all the stand-
ards (BHET, MHET, TPA) were prepared in 100% metha-
Purification of the Obtained TPA nol. Mobile phases containing acetonitrile, 10 mM sulphuric
acid, and ultrapure water was used with a UV detector of
Purification of the TPA precipitated from the monomers’ wavelength 254 nm. The total run time for each sample was
mixture was carried out using a dissolution/crystallisation 18 min. The quantification of the produced monomers was
method with slight modifications [31]. For this purpose, 1 g carried out by plotting calibration curves within the range
of the precipitated TPA was dissolved in dimethyl acetamide of (0.01–1 mM) standard concentrations.
(DMAc) at its boiling point. The hot solution was then fil- NMR measurements of samples were carried out by a
tered with filter paper to remove the insoluble parts from the Bruker NMR advanced using the operating frequency of
solution system. The filtrate was directly inserted into an ice 600 MHz at ­25°C in d-DMSO.
bath, and the formed TPA crystals were collected by filtra- ICP-MS measurements were performed with Agilent
tion, washed with excess amount of ethanol to get rid of the 7800 ICP-MS instrument equipped with an SPS4 autosa-
DMAc residuals and finally dried in an oven at 80°C for 24 h. mpler, MS40 + Foreline Pump, and G8481A re-circulating
chiller. As the sample preparation for the test, 0.1 g of the
Instrumental Characterisations sample was added to a beaker and digested with 5 mL of
67% Nitric Acid, 1.7 mL concentrated HCl, and 3.3 mL of
Fourier transform infrared spectroscopy (FTIR) (Perkin water. The digestion process was carried out in a microwave
Elmer, UK) was used to analyse PET samples before and at ­23°C for 20 min. Finally, the prepared sample was rinsed
after microwave treatments and produced monomers within and made up to 50 mL with water for ICP-MS analysis.
spectral range of 4000–600 ­cm−1 accumulating 16 scan The acid number values of the products were determined
cycles and a resolution of 4 ­cm−1. The carbonyl index was with Eutech ION 700 pH/mV meter by following the ASTM
extracted from obtained data by taking ratios of intensities standard D8031-16, the standard test method for the acid
of ester carbonyl peak (1713 ­cm−1) to normal C-H bonding number of terephthalic acid by automatic potentiometric
mode (1408 ­cm−1) by using following equation [32, 33] titration.
Absorption at 1713cm−1 APHA colour values were determined based on a colour
Carbonyl index = (3) test by using Varian Cary 50 UV–Vis Spectrometer. As the
Absorption at 1408cm−1
sample preparation, 2.5 g of the sample was dissolved in

13
Journal of Polymers and the Environment

50 g of DMF (%5) by assisting with an ultrasonic bath. The one. As demonstrated in Table 2, The coefficient of deter-
measurements were performed with UV–Vis Calibrated at mination of all the studied responses were within acceptable
a low Range of 10–200 Pt–Co and a high Range of 50–500 ranges and the obtained adjusted ­R2 values were 0.9632,
Pt–Co. Samples were analysed by an appropriate calibra- 0.9487, and 0.9206 for carbonyl index, weight loss at T ­ 0 of
tion curve. degradation and crystallinity index respectively. High model
F-values were also observed which implied that the quad-
ratic model was suitable for the prediction of the studied
responses. Analysis of variance (ANOVA) was performed to
Results and Discussion
evaluate the curvature significance in the responses at a 95%
confidence interval and showed that all the responses’ prob-
MW Assisted Pre‑treatment of PET
ability values (p-value) were less than 0.05. The lack of fit
test also demonstrated a highly desirable nonsignificant lack
In this study, green microwave technology was employed
of fit (p > 0.1) where the carbonyl index, weight loss at T ­0
for PET pre-treatment using fully biobased, inexpensive
of degradation and crystallinity index had p-values of 0.81,
and nontoxic glycerol solvent. Based on our previous work,
0.74, and 0.90 respectively. Detailed results of the ANOVA
a significant catalytic activity and increased reaction rate
for the response surface quadratic model and the significance
under MW irradiations was observed in solvents which pos-
of its regression coefficients are illustrated in Tables S2, S3
sess high pH and low density values [10, 16] . For the glyc-
and S4 for the carbonyl index, weight loss at T ­ 0 of degra-
erol solvent system, pH and density values were observed
dation and crystallinity index, respectively. The analysis of
to be 8.6 and 1.21 g/cm3 respectively [35]. The significant
normal distribution of residual values which explain the dif-
effect of glycerol solvent in its interaction with PET lies in
ference between predicted values (model) and the observed
its ability to raise the electron density and attraction between
values (experimental) is another critical diagnostic tool for
carbonyl oxygen of PET which tend to form single bond
evaluating the suitability of the fitted model to predict the
after rearranging electrons. Thus, electrostatic attraction
studied responses. The data is said to be distributed normally
between H and O can give glycol and forms ester C=O again
if all the points fall close to the straight line. The normal
which facilitate producing monomers and other intermedi-
probability plots of residual values of the carbonyl index,
ate products with leftover glycol as presented in Fig. 1 [36].
weight loss at ­T0 of degradation and crystallinity index are
illustrated in Fig. S1. In each plot, the residual values for
Experimental Design for MW‑Assisted PET most of performed experiments were evenly distributed
Pre‑treatment in space. Thus, the model can be accepted to describe the
responses under study.
The response surface methodology was carried out to evalu- The relationship between the studied responses and
ate MW-assisted pre-treatment of PET best fitting model. In parameters of the MW assisted pre-treatment can be also
the current work, based on the BBD design for the studied demonstrated through the coefficient of determination for
parameters (MW irradiation time, volume of glycerol, and the quadratic equations as shown in Table 2. The carbonyl
MW power), 15 experimental runs were performed. After index ­(Y1) of all pre-treated PET was increased when com-
each experiment, the crystallinity and carbonyl indices of pared to that of virgin untreated PET (2.80). Both the MW
PET residues, and weight loss % of PET at ­T0 of degradation power and irradiation time had a negative effect on the car-
were determined and presented in Table S1. bonyl index, whereas volume of glycerol had a significant
The responses values obtained from the MW assisted pre- positive effect. The interactions of both glycerol volume and
treatment of PET were tested against various statistical models irradiation time with MW power indicated a negative effect
and the quadratic model was found to be the best fitting on pre-treated PET carbonyl index values. Thus, it can be

Fig. 1  Proposed interaction


mechanism between glycerol
and PET in the MW assisted
pre-treatment step

13
Journal of Polymers and the Environment

Table 2  Statistical analysis of Fitting model Factors Coefficient P-value ANOVA


the studied responses for the
MW assisted PET pre-treatment PET Carbonyl index (­ Y1) Intercept 3.30 F = 14.54,
X1 − 0.0313 0.4934 R2 = 0.9632, Model P-value 0.0033,
P-value of lack of fit = 0.8112
X2 − 0.1388 0.0220
X3 0.2775 0.0012
X1X2 − 0.0075 0.9052
X1X3 0.1950 0.0225
X2X3 − 0.1500 0.050
X12 − 0.1396 0.0751
X22 0.2754 0.0069
X32 0.3679 0.0020
PET Crystallinity index ­(Y2) Intercept 57.54 F = 10.26,
X1 0.5100 0.050 R2 = 0.9487, Model P-value
= 0.0098,
X2 6.24 0.0030
P-value of lack of fit = 0.9039
X3 2.63 0.001
X1X2 7.35 0.0065
X1X3 3.61 0.0092
X2X3 0.8925 0.0001
X12 − 5.62 0.0217
X22 − 8.37 0.0045
X32 − 2.17 0.2597
­ 0 of
Weight loss of PET at T Intercept 9.97 F = 6.44,
degradation ­(Y3) X1 0.1561 0.001 R2 = 0.9206, Model P-value
= 0.0270,
X2 2.84 0.0026
P-value of lack of fit = 0.7475
X3 − 0.5352 0.3412
X1X2 1.77 0.0576
X1X3 − 1.04 0.2065
X2X3 − 0.8040 0.3149
X12 0.7810 0.3450
X22 − 2.72 0.0152
X32 − 1.01 0.2343

X1: MW irradiation time, ­X2: MW power, ­X3: volume of glycerol

predicted that high glycerol volumes with low MW power Graphical plots of surface response analysis were gen-
and irradiation time lead to an increase in carbonyl groups erated between the studied variables, and responses were
due to surface oxidation of the treated PET [37]. obtained to determine their effect in combinations. The 3D
Moreover, the quadratic model equation coefficients and contour plots for the, carbonyl index, crystallinity index
showed an increase in crystallinity index of treated PET ­(Y2) and weight loss of PET at T ­ 0 of degradation are shown in
in all studied responses and their interactions. The preferen- Figs. 2, 3, 4, respectively while the perturbation plots repre-
tial degradation of amorphous part of semicrystalline PET senting the combinatorial effects of all the studied factors on
gave rise to an increased crystallinity index when compared each response are found in Fig. S2. The Carbonyl index is a
to PET virgin counterpart (31.4%) at high levels of the stud- critical parameter for evaluating the extent of surface modi-
ied factors [10, 38]. Positive effects were observed with MW fication of polymers after treatments. Figure 2 illustrates the
power and irradiation time along with their interactions on carbonyl index ­(Y1) and its dependence on MW time, power,
PET weight loss (­ Y3) while glycerol volume and its inter- and glycerol volume. It can be noticed that the interactions
actions with MW power and irradiation time had negative of high volumes of glycerol with low levels of MW power
effect. This indicated that high levels of irradiation times and and time showed a significant rise in carbonyl index values
MW power coupled with low glycerol volume could lower up to 4.43 which confirmed the oxidation of the PET surface
the thermal stability of PET and facilitate initial depolymeri- at these levels rather than complete depolymerisation [39].
sation of PET at its ­T0 of degradation. As demonstrated in counter and 3D plots for crystal-
linity index (­ Y2) in Fig. 3, high levels of MW power and

13
Journal of Polymers and the Environment

Fig. 2  3D and contour plots of the effect of the interaction of a power ­(X2) and glycerol volume ­(X3), b time ­(X1) and power ­(X2) and c glycerol
volume ­(X3) and time (­ X1) on carbonyl index

Fig. 3  3D and contour plots of the effect of the interaction of a time (­X1) and power (­X2), b power ­(X2) and glycerol volume (­X3), and c time
­(X1) and glycerol volume ­(X3) on the crystallinity index

13
Journal of Polymers and the Environment

Fig. 4  3D and contour plots of the effect of the interaction of a time ­(X1) and power ­(X2), b time ­(X1) and glycerol volume ­(X3) and c power
­(X2) and glycerol volume ­(X3) on PET weight loss (%)

time led to the increase of crystallinity index of pre-treated of degradation that reached up to 11.9% while pure PET
PET. This result could be an indication of the initial deg- weight loss had a value of 0.44%.
radation of the amorphous part of PET, which caused an
increase in its overall crystallinity [32, 40]. Moreover, Optimisation of MW‑Assisted Pre‑treatment of PET
high values of crystallinity index were also observed in
large volumes of glycerol up to a certain level, indicating The BBD optimisation methodology was conducted in
that the intermediate levels of glycerol are relatively more order to obtain maximum PET weight loss and desirable
effective for the pre-treatment of PET. characteristics of the pre-treated PET to accelerate the
The PET weight loss (%) at T ­ 0 ­( Y 3) is also an influ- PET depolymerisation in the second step of MW assisted
ential variable for evaluating the initial degradation of hydrolysis. The MW assisted pre-treatment conditions
PET through its thermal stability. High percentages of were adjusted in such a way to get maximum weight loss
PET weight loss demonstrate a decreased level of ther- (%) at T­ 0, maximum carbonyl index and minimum crystal-
mal stability which validates initial degradation of PET. linity index as shown in Table 3. A batch experiment was
As shown in Fig. 4, the interaction between MW power conducted by mixing 1 gm of PET with 12 mL of glycerol
and time has positive effect on PET weight loss (%) at ­T0 and exposing the mixture for 3 min to MW irradiation

Table 3  The observed and predicted response values of optimised MW assisted pre-treatment of PET
Independent variable Optimised level

X1: MW time (min) 3.0


X2: MW power (W) 182
X3: Volume of glycerol (mL) 12.0
Over all desirability 0.79
Dependent variables Desirability Expected Observed

Y1: PET carbonyl index Maximise 4.43 4.22


Y2: PET crystallinity index Minimise 35.45 32.98
Y3: PET weight loss at T­ 0 (%) Maximise 9.20 10.25

13
Journal of Polymers and the Environment

of 182W power. The predicted values for the optimised TPA was increased from 12% for virgin PET to 79.1% in
conditions showed fine agreement with the observed ones the pre-treated PET confirming the enhanced hydrolytic
and that indicated a successful optimisation of the proposed depolymerisation of PET upon applying the MW assisted
PET pre-treatment process. The optimised pre-treated PET pre-treatment step.
had a 10.25% degradation % at T ­ 0, an elevated carbonyl
index of 4.22 and 33% crystallinity. Purification and Validation of Obtained TPA
for Repolymerisation

MW Assisted‑Hydrolysis of PET A dissolution/recrystallisation purification approach was


applied on the TPA produced from the pre-treated PET
Based on the optimised conditions of PET MW assisted depolymerisation to remove any traces of monomers or
pre-treatment, both virgin PET and pre-treated PET were impurities from it. The TPA before and after purification
hydrolysed under MW irradiations using N ­ aHCO3 in EG step were termed as precipitated TPA and recrystallised
for 3 min. After the MW treatment, pre-treated PET was TPA respectively. As a result of the applied purification
separated and PET conversion (%) was calculated. The methodology, 72% of recrystallised TPA was recovered
produced monomers in the filtrate were also analysed by from the precipitated monomer mixture. This decreased
HPLC and the effects of various depolymerisation steps recovery value could be attributed to the nature of purifica-
on selectivity of monomers (BHET, MHET), yield of tion via crystallisation since an amount of the desired com-
TPA were determined as shown in Fig. S3, and Fig. 5. pound might remain in the solvent system at the end of the
The observed conversion value of PET was only 61% after heating–cooling process. Such drawback can be resolved
applying hydrolysis reaction on untreated virgin PET. On by cooling the solvent system at lower temperatures. The
the other hand, a 99.9% PET conversion was achieved recrystallised TPA purity was then confirmed using different
when the same hydrolysis reaction was applied to pre- characterisation techniques.
treated PET obtained from the MW assisted pre-treatment
step. Such improved efficiency of the hydrolysis reaction FTIR
on the pre-treated PET could be associated to the struc-
tural modifications and new properties that were developed The identity of the precipitated TPA monomer before and
in such residual. Concerning the produced monomers anal- after purification was evaluated using FTIR. As illustrated
ysis, a higher selectivity of MHET was observed as com- in Fig. S4, the TPA obtained from PET depolymerisation
pared to BHET and TPA in hydrolysis step of untreated showed similar structural characteristics to that of standard
PET while TPA was the main monomer produced from TPA. The peaks at frequencies 3064 ­cm−1, 1673 ­cm−1, and
the pre-treated PET residue. Additionally, the yield of 1280 ­cm−1 corresponded to carboxylic acid –OH, carbonyl
group, and ether C–O stretching respectively. For the precip-
itated TPA, the extra peaks appearing in the fingerprint area
which might be related to the MHET and BHET monomers
present as impurities in the precipitated powder were suc-
cessfully removed in the crystallised TPA. Thus, indicating
its acceptable purity.

NMR

The structural confirmation of the TPA before and after puri-


fication was performed by NMR spectroscopy. The 1H and
13
C NMR spectra of the precipitated TPA product after PET
depolymerization are presented in Fig. 6A and C, respec-
tively. The signs labelled as “a” and “b” on the 1H NMR
spectrum were assigned to acidic –OH, and –CH groups
on the benzene rings at 13.34 (s, 2H) ppm and 8.09 (s, 4H),
respectively. In addition to the NMR signals belonging to
the TPA monomer, the signals assigned as “aI” and “bI”,
corresponding to the protons of the aromatic ring and acidic
Fig. 5  The percentage conversion of PET with different treatment
methodologies, and the selectivity of obtained monomers (BHET, –OH, were attributed to the MHET monomer. Moreover, the
MHET) and the yield of TPA signals arising from the MHET monomer (labelled as “c”

13
Journal of Polymers and the Environment

Fig. 6  1H NMR spectrum of a Precipitated TPA and b crystallised TPA and 13C NMR of c Precipitated TPA and d crystallised TPA

and “d”) appeared on the spectrum, corresponding to meth- spectrum (Fig. 6D), having the signals at 166.63 134.40,
ylene linked to the –OH groups (δH = 3.71 ppm, m, 4H) and and 129.41 ppm and confirming that the TPA monomer was
methylene linked to the -COO groups (δH = 4.32 ppm, t,4H), successfully extracted from the monomer mixture. These
respectively. As to the signs around 2.5 ppm, these belonged results also came in coordination with HPLC analysis of
to DMSO, while the ones at 3.3 and 2.0 ppm were referred the crystallised TPA powder where a TPA yield of 94.88%
to water residual and solvent contamination. As shown in was obtained.
Fig. 6C, the C NMR signals belonging to the TPA monomer
are labelled as 1 (δC = 166.57 ppm), 2 (δC = 134.27 ppm), 3
(δC = 129.44 ppm) while the signals belonging to the MHET APHA Colour Value
monomer and labelled as 1­ I and 1­ II (δC = 165.22 ppm and
165.47 ppm), ­2I (δC = 133.34 ppm), ­3I (δC = 129.46 ppm), The colour of TPA is one of the important physical param-
4 (δC = 66.92 ppm) and 5 (δC = 58.90 ppm) were assigned eters that should be considered for the re-polymerization
to the carbons of the chemical structure of both monomers step due to the fact that this colour is transferable to the
[31, 41]. PET product making it undesirable for industrial manufac-
After the purification process, it was demonstrated that turing. The colour of the obtained TPA was evaluated using
the crystallised TPA monomer was successfully separated a spectrophotometric methodology and its APHA colour
from the monomers’ mixture giving rise to pure peaks in value was determined [42]. As shown in Fig. 7a, the APHA
both 1H and C NMR spectra. As presented in Fig. 6B, the value of the standard TPA was calculated as 7.71, while the
1
H NMR signs, 13.28 (s, 2H) ppm and 8.07 (s, 4H), cor- precipitated TPA from PET depolymerization process had
responding to the TPA monomer structure were the most APHA value of 245.2. This significantly high APHA colour
prominent signals remaining on the spectrum after the puri- value could be related to the contamination of the precipi-
fication of the TPA monomer. Similarly, only peaks belong- tated TPA by other organic compounds that were produced
ing to the TPA structure were observed on the 13C NMR during the PET depolymerisation process. Noticeably, after

13
Journal of Polymers and the Environment

Fig. 7  APHA colour values (a) and acid number (b) of the standard TPA, and the depolymerised product before and after the crystallisation of
the TPA monomer

the application of the dissolution and crystallisation pro- Table 4  Heavy metal trace analysis of the samples
cedure on the precipitated TPA, these organic compounds Heavy metal Samples
were removed and a crystallised TPA of 82.4 APHA value (µg/g)
was obtained. Standard TPA Precipitated crystallised TPA
TPA

Aluminium 1.3 3.0 1.6


Acid Number Value
Titanium < 0.5 < 0.5 < 0.5
Chromium < 0.5 < 0.5 < 0.5
The acid number is defined as a number utilised to quan-
Manganese < 0.5 < 0.5 < 0.5
tify the acidity of a chemical substance which is critical for
Cobalt < 0.5 < 0.5 < 0.5
the re-polymerisation of the TPA monomer in our case. As
Molybdenum < 0.5 < 0.5 < 0.5
shown in Fig. 7b, standard TPA has an acid number of 674.7
Antimony < 0.5 1.1 0.8
mgKOH/g, while the precipitated TPA post PET depolymer-
Iron 1.1 1.7 1.3
isation had an acid value of 329.3 mgKOH/g. Such signifi-
cantly low acid number could be attributed to the presence
of MHET monomer and other residuals in the precipitated
metallic content than that of the standard TPA. On the other
product resulting in the significantly low total acidity of
hand, the crystallised TPA possessed almost similar heavy
the material. Upon purification of the precipitated TPA
metal content in comparison with the standard TPA, thus,
monomer, the acid number value was increased to 645.3
making it suitable for the production of PET.
mgKOH/g, giving rise to a TPA monomer having sufficient
colour and acid number values for re-polymerisation into
PET material.
Conclusion

Heavy Metal Content In the current study, a green, straight-forward 2 step 100%
PET depolymerisation process with high yields of purified
The heavy metal content of the TPA, which is the main mon- TPA monomer utilising low energy, MW technology and
omer used for producing PET, should be within an accept- green solvents is presented. The proposed depolymerisa-
able range to avoid contaminating the PET product and to tion methodology completely depolymerised PET into its
keep the polymerisation reaction efficient. Bearing this in valuable monomers within a total of 6 min under environ-
mind, the precipitated TPA, before and after purification mentally friendly and mild operating conditions. PET was
was analysed by ICP-MS to determine its heavy metal con- initially pre-treated via a green MW method using glycerol
tent. As listed in Table 4, it was found that the precipitated as a treating solvent to render PET more amenable for com-
TPA post the depolymerisation process had a slightly higher plete degradation. The pre-treated PET residue was further

13
Journal of Polymers and the Environment

depolymerised via a mild hydrolytic MW assisted process 4. Chan K, Zinchenko A (2021) Conversion of waste bottles’ PET to
using ­NaHCO3 in EG. A precipitated TPA yield of 79% was a hydrogel adsorbent via PET aminolysis. J Environ Chem Eng.
https://​doi.​org/​10.​1016/j.​jece.​2021.​106129
confirmed using HPLC analysis. Furthermore, taking into 5. Pudack C, Stepanski M, Fässler P (2020) PET recycling – con-
the consideration the importance of TPA purity for repoly- tributions of crystallization to sustainability. Chem-Ing-Tech
merisation into industrially acceptable PET, a purification 92:452–458. https://​doi.​org/​10.​1002/​cite.​20190​0085
process for the precipitated TPA was applied using a dis- 6. Kirshanov K, Toms R, Aliev G et al (2022) Recent develop-
ments and perspectives of recycled poly(ethylene terephthalate)-
solution/recrystallisation approach. A TPA purity of 95%, based membranes: a review. Membranes (Basel). https://​doi.​org/​
82.4 APHA colour value, 645.3 mgKOH/g acid number 10.​3390/​membr​anes1​21111​05
and acceptable heavy metal content was achieved, making 7. Sadeghi B, Marfavi Y, AliAkbari R et al (2021) Recent studies
the obtained TPA comparable with commercial standard on recycled PET fibers: production and applications: a review.
Mater Circ Econ. https://​doi.​org/​10.​1007/​s42824-​020-​00014-y
equivalent. 8. Miller KK (2015) Process for preparation of terephthalic acid.
Michigan State University, East Lansing
Supplementary Information The online version contains supplemen- 9. Frisa-Rubio A, González-Niño C, Royo P et al (2021) Chemi-
tary material available at https://d​ oi.o​ rg/1​ 0.1​ 007/s​ 10924-0​ 23-0​ 2979-8. cal recycling of plastics assisted by microwave multi-frequency
heating. Clean Eng Technol. https://​doi.​org/​10.​1016/j.​clet.​2021.​
Acknowledgements This project received funding from the Euro- 100297
pean Union’s Horizon 2020 research, the Irish Research Council 10. Azeem M, Fournet MB, Attallah OA (2022) Ultrafast 99% Poly-
(GOIPG/2021/1739) and innovation programme under grant agreement ethylene terephthalate depolymerization into value added mono-
No. 870292 (BIOICEP) and was supported by the National Natural mers using sequential glycolysis-hydrolysis under microwave
Science Foundation of China (grant numbers: Institute of Microbiol- irradiation. Arab J Chem 15:103903. https://​doi.​org/​10.​1016/j.​
ogy, Chinese Academy of Sciences: 31961133016; Beijing Institute arabjc.​2022.​103903
of Technology: 31961133015; Shandong University: 31961133014). 11. Gómez AV, Biswas A, Tadini CC et al (2019) Use of natural
deep eutectic solvents for polymerization and polymer reactions.
Author Contributions “Conceptualization, MA; methodology, MA, J Braz Chem Soc 30:717–726. https://​doi.​org/​10.​21577/​0103-​
CET, and OA software, MA; validation, MA, OA, CET, and MBF; 5053.​20190​001
formal analysis, MA, OA, CET, and MBF; investigation, MA, CET, 12. Sheel A, Pant D (2019) 4 - Chemical Depolymerization of PET
and OA; resources, MBF; data curation, MA, CET, and OA; writ- Bottles via Glycolysis, In: Recycling of Polyethylene Tereph-
ing—original draft preparation, MA; writing—review and editing, MA, thalate Bottles, Thomas J, Rane A, Kanny K, VK A, Thomas
OA, CET, and MBF; visualization, MA and OA; supervision, MBF, MG (Eds) William Andrew Publishing, pp 61–84. https://​doi.​
OA, and CET; project administration, MBF; funding acquisition, MA org/​10.​1016/​B978-0-​12-​811361-​5.​00004-3
and MBF. All authors have read and agreed to the published version 13. Tang X, Chen EYX (2019) Toward infinitely recyclable plastics
of the manuscript.” derived from renewable cyclic esters. Chem 5:284–312. https://​
doi.​org/​10.​1016/j.​chempr.​2018.​10.​011
Funding This study was supported by Irish Research Council, 14. Liu B, Fu W, Lu X et al (2019) Lewis acid-base synergis-
GOIPG/2021/1739. tic catalysis for polyethylene terephthalate degradation by
1,3-dimethylurea/Zn(OAc)2 deep eutectic solvent. ACS Sustain
Data Availability All data generated or analysed during this study are Chem Eng 7:3292–3300. https://​doi.​org/​10.​1021/​acssu​schem​
included in the article. eng.​8b053​24
15. Cot S, Leu MK, Kalamiotis A et al (2019) An oxalate-bridged
Declarations binuclear iron(III) ionic liquid for the highly efficient glycolysis of
polyethylene terephthalate under microwave irradiation. ChemP-
Conflict of interest There are no conflicts to declare. lusChem 84:786–793. https://​doi.​org/​10.​1002/​cplu.​20190​0075
16. Attallah OA, Janssens A, Azeem M, Fournet MB (2021) Fast, high
monomer yield from post-consumer polyethylene terephthalate via
combined microwave and deep eutectic solvent hydrolytic depo-
lymerization. ACS Sustain Chem Eng. https://​doi.​org/​10.​1021/​
References acssu​schem​eng.​1c071​59
17. Xi G, Lu M, Sun C (2005) Study on depolymerization of waste
1. Benyathiar P, Kumar P, Carpenter G et al (2022) Polyethylene polyethylene terephthalate into monomer of bis(2-hydroxyethyl
terephthalate (PET) bottle-to-bottle recycling for the beverage terephthalate). Polym Degrad Stab 87:117–120. https://​doi.​org/​
industry: a review. Polymers (Basel). https://​doi.​org/​10.​3390/​ 10.​1016/j.​polym​degra​dstab.​2004.​07.​017
polym​14122​366 18. Shukla SR, Harad AM (2005) Glycolysis of polyethylene tereph-
2. Padhan RK, Sreeram A (2019) 7 - Chemical Depolymerization thalate waste fibers. J Appl Polym Sci 97:513–517. https://d​ oi.o​ rg/​
of PET Bottles via Combined Chemolysis Methods, In: Recycling 10.​1002/​app.​21769
of Polyethylene Terephthalate Bottles, Thomas S, Rane A, Kanny 19. Fiore V, Scalici T, Nicoletti F et al (2016) A new eco-friendly
K, VK A, Thomas MG, (eds) William Andrew Publishing, pp chemical treatment of natural fi bres: effect of sodium bicarbonate
135–147. https://​doi.​org/​10.​1016/​B978-0-​12-​811361-​5.​00007-9 on properties of sisal fi bre and its epoxy composites. Compos Part
3. Thiyagarajan S, Maaskant-Reilink E, Ewing TA et al (2022) Back- B 85:150–160. https://d​ oi.o​ rg/1​ 0.1​ 016/j.c​ ompos​ itesb.2​ 015.0​ 9.0​ 28
to-monomer recycling of polycondensation polymers: opportuni- 20. Ezeanu DS, Matei D (2021) Natural depolymerization of waste
ties for chemicals and enzymes. RSC Adv 12:947–970. https://d​ oi.​ poly (ethylene terephthalate) by neutral hydrolysis in marine
org/​10.​1039/​d1ra0​8217e water. Sci Rep. https://​doi.​org/​10.​1038/​s41598-​021-​83659-2
21. Sun CH, Chen XP, Zhuo Q, Zhou T (2018) Recycling and depo-
lymerization of waste polyethylene terephthalate bottles by

13
Journal of Polymers and the Environment

alcohol alkali hydrolysis. J Cent South Univ 25:543–549. https://​ 33. Chelliah A, Subramaniam M, Gupta R, Gupta A (2017) Evalu-
doi.​org/​10.​1007/​s11771-​018-​3759-y ation on the thermo-oxidative degradation of PET using pro-
22. Pagliaro M, Ciriminna R, Kimura H et al (2007) From glycerol degradant additives. Indian J Sci Technol 10:1–5. https://​doi.​org/​
to value-added products. Angew Chemie - Int Ed 46:4434–4440. 10.​17485/​ijst/​2017/​v10i6/​111212
https://​doi.​org/​10.​1002/​anie.​20060​4694 34. Li W, Zhang C, Chi H et al (2017) Development of antimicrobial
23. Dzhabarov G, Sapunov V, Kozlovskiy R et al (2020) A method of packaging film made from poly(lactic acid) incorporating titanium
polyethylene terephthalate depolymerization by biodiesel wastes. dioxide and silver nanoparticles. Molecules. https://​doi.​org/​10.​
Pet Coal 62:19–26 3390/​molec​ules2​20711​70
24. Choi S, Choi HM (2019) Eco-friendly, expeditious depolym- 35. Wernke MJ (2014) Glycerol. Encycl Toxicol Third Ed. https://d​ oi.​
erization of PET in the blend fabrics by using a bio-based deep org/​10.​1016/​B978-0-​12-​386454-​3.​00510-8
eutectic solvent under microwave irradiation for composition 36. Carta D, Cao G, D’Angeli C (2003) Chemical recycling of
identification. Fibers Polym 20:752–759. https://​doi.​org/​10.​1007/​ poly(ethylene terephthalate) (PET) by hydrolysis and glycolysis.
s12221-​019-​8931-y Environ Sci Pollut Res 10:390–394. https://d​ oi.o​ rg/1​ 0.1​ 065/e​ spr2​
25. González-Rivera J, Husanu E, Mero A et al (2020) Insights into 001.​12.​104.8
microwave heating response and thermal decomposition behavior 37. Rouillon C, Bussiere PO, Desnoux E et al (2016) Is carbonyl index
of deep eutectic solvents. J Mol Liq. https://d​ oi.o​ rg/1​ 0.1​ 016/j.m
​ ol- a quantitative probe to monitor polypropylene photodegradation?
liq.​2019.​112357 Polym Degrad Stab 128:200–208. https://d​ oi.o​ rg/1​ 0.1​ 016/j.p​ olym​
26. González-Rivera J, Mero A, Husanu E et al (2021) Combining degra​dstab.​2015.​12.​011
acid-based deep eutectic solvents and microwave irradiation 38. Beltrán-Sanahuja A, Casado-Coy N, Simó-Cabrera L, Sanz-
for improved chestnut shell waste valorization. Green Chem Lázaro C (2020) Monitoring polymer degradation under different
23:10101–10115. https://​doi.​org/​10.​1039/​d1gc0​3450b conditions in the marine environment. Environ Pollut. https://​doi.​
27. Zhang S, Xu W, Du R et al (2022) Cosolvent-promoted selective org/​10.​1016/j.​envpol.​2019.​113836
non-aqueous hydrolysis of PET wastes and facile product separa- 39. Appu SP, De SK, Khan MJ, Al-Harthi MA (2013) Natural weather
tion. Green Chem. https://​doi.​org/​10.​1039/​d2gc0​0328g ageing of starch/polyvinyl alcohol blend: Effect of glycerol content. J
28. Hu LC, Oku A, Yamada E, Tomari K (1997) Alkali-decomposi- Polym Eng 33:257–263. https://​doi.​org/​10.​1515/​polye​ng-​2012-​0048
tion of poly(ethylene terephthalate) in mixed media of nonaque- 40. Choi HM, Cho JY (2016) Microwave-mediated rapid tailoring of
ous alcohol and ether. Study on recycling of poly(ethylene tere- PET fabric surface by using environmentally-benign, biodegrad-
phthalate). Polym J 29:708–712. https://​doi.​org/​10.​1295/​polymj.​ able urea-choline chloride deep eutectic solvent. Fibers Polym
29.​708 17:847–856. https://​doi.​org/​10.​1007/​s12221-​016-​5334-1
29. Hu L-C, Oku A, Yamada E, Tomari K (1997) Alkali-decompo- 41. Lima GR, Monteiro WF, Ligabue R, Santana RMC (2017) Titan-
sition of poly (ethylene terephthalate) in mixed media of non- ate nanotubes as new nanostrutured catalyst for depolymerization
aqueous alcohol and ether. Study on recycling of poly (ethylene of PET by glycolysis reaction. Mater Res 20:596–602. https://d​ oi.​
terephthalate). Polym J 29:708–712 org/​10.​1590/​1980-​5373-​mr-​2017-​0645
30. Du JT, Sun Q, Zeng XF et al (2020) ZnO nanodispersion as pseu- 42. Taheri M, Jahanfar M, Ogino K (2019) Synthesis of acrylic resins
dohomogeneous catalyst for alcoholysis of polyethylene tereph- for high-solids traffic marking paint by solution polymerization.
thalate. Chem Eng Sci 220:115642. https://​doi.​org/​10.​1016/j.​ces.​ Des Monomers Polym 22:213–225. https://d​ oi.o​ rg/1​ 0.1​ 080/1​ 5685​
2020.​115642 551.​2019.​16993​49
31. Wu SC, Cheng ZM, Wang SD, Shan XL (2011) Recovery of tere-
phthalic acid from alkali reduction wastewater by cooling crys- Publisher's Note Springer Nature remains neutral with regard to
tallization. Chem Eng Technol 34:1614–1618. https://​doi.​org/​10.​ jurisdictional claims in published maps and institutional affiliations.
1002/​ceat.​20110​0096
32. Attallah OA, Azeem M, Nikolaivits E et al (2022) Progress- Springer Nature or its licensor (e.g. a society or other partner) holds
ing ultragreen, energy-efficient biobased depolymerization of exclusive rights to this article under a publishing agreement with the
poly(Ethylene terephthalate) via microwave-assisted green deep author(s) or other rightsholder(s); author self-archiving of the accepted
eutectic solvent and enzymatic treatment. Polymers (Basel) manuscript version of this article is solely governed by the terms of
14:109 such publishing agreement and applicable law.

Authors and Affiliations

Muhammad Azeem1 · Olivia A. Attallah1,2 · Cuneyt Erdinc Tas1 · Margaret Brennan Fournet1

2
* Muhammad Azeem Pharmaceutical Chemistry Department, Faculty of Pharmacy,
azeemm1047@gmail.com Heliopolis University, Cairo ‑ Belbeis Desert Road, El Salam,
Cairo 11777, Egypt
1
Centre for Polymer Sustainability, PRISM Research
Institute, Technological University of the Shannon: Midlands
Midwest, Athlone N37 HD68, Ireland

13

You might also like