Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

acssuschemeng.3c03114

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/372419807

Catalytic Fabric Recycling: Glycolysis of Blended PET with Carbon Dioxide and
Ammonia

Article in ACS Sustainable Chemistry & Engineering · July 2023


DOI: 10.1021/acssuschemeng.3c03114

CITATIONS READS

0 36

10 authors, including:

Yang Yang Shriaya Sharma


University of Copenhagen University of Copenhagen
3 PUBLICATIONS 14 CITATIONS 1 PUBLICATION 0 CITATIONS

SEE PROFILE SEE PROFILE

Carlo Di Bernardo Elisa Rossi


University of Copenhagen Università di Pisa
1 PUBLICATION 0 CITATIONS 8 PUBLICATIONS 22 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

R E V I E W Proteomic Applications in Antimicrobial Resistance and Clinical Microbiology Studies View project

Proteomic Applications in Antimicrobial Resistance and Clinical Microbiology Studies View project

All content following this page was uploaded by Carlo Di Bernardo on 18 July 2023.

The user has requested enhancement of the downloaded file.


pubs.acs.org/journal/ascecg Research Article

Catalytic Fabric Recycling: Glycolysis of Blended PET with Carbon


Dioxide and Ammonia
Yang Yang, Shriaya Sharma, Carlo Di Bernardo, Elisa Rossi, Rodrigo Lima, Fadhil S. Kamounah,
Margarita Poderyte, Kasper Enemark-Rasmussen, Gianluca Ciancaleoni,* and Ji-Woong Lee*
Cite This: https://doi.org/10.1021/acssuschemeng.3c03114 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF NAPLES FEDERICO II on July 18, 2023 at 09:36:06 (UTC).

ABSTRACT: The ubiquity of nonbiodegradable polyethylene terephthalate (PET) materials has led to significant waste
management challenges. Although PET plastics can be recycled, blended materials, such as PET/cotton fabrics, complicate the
recycling process due to the labile glycosidic bonds in cotton. In this study, we present a practical and scalable approach for recycling
of PET and PET/cotton interwoven fabrics via catalytic glycolysis with ammonium bicarbonate (NH4HCO3), which decomposed to
ammonia, carbon dioxide, and water. This catalytic approach outperformed conventional acid/base and metal catalysis in selectively
recovering and upcycling cotton-based materials. We demonstrated the large-scale recovery of textile from blended fabrics (up to
213 g), showcasing the advantages of traceless catalysis using ammonia and CO2 from ammonium bicarbonate. Owing to our metal-
free reaction conditions, high-purity bis(hydroxyethyl)terephthalate (BHET) was obtained which was thermally repolymerized to
PET. Through thermal analysis, kinetics, and control experiments, we show that ammonia and CO2 are crucial for achieving optimal
glycolysis via transesterification. Our method offered a traceless, environmentally friendly, and practical approach for polyester
recycling and cotton recovery, representing a significant step toward sustainable, closed-loop production of plastics and textiles.
KEYWORDS: polyethylene terephthalates, plastic, fabric, carbon dioxide, glycolysis

■ INTRODUCTION
Synthetic polymers represent one of the most significant
being either landfilled or incinerated.11 Therefore, recycling
these composite-plastic-based materials has become an urgent
polluting agents threatening the sustainable growth of human societal issue. Chemical recycling or upcycling of plastic waste
society. The worldwide polymer production (ca. 388 million has emerged as an attractive option to mitigate fossil fuel
metric tons in 2019) has grown exponentially over the years.1,2 depletion, reduce greenhouse gas emissions, and minimize
Among them, polyethylene terephthalate (PET) stands out as negative environmental impacts resulting from energy recovery
one of the most commonly produced polymers, with ca. 70 in waste incinerators.12−16
million metric tons being produced every year and an expected The catalytic and enzymatic glycolysis of PET yields the
annual growth rate exceeding 4%.3,4 PET-based polymers are valuable intermediate bis(hydroxyethyl)terephthalate (BHET)
primarily utilized in the production of beverage bottles, while (Figure 1A,B),17−19 which can be recycled as a precursor for
approximately one-third of the produced PET is blended with PET via repolymerization20−25 and other biodegradable
cotton-based fibers and viscose to manufacture fabrics in the
textile industry.5,6 In 2020 alone, over 109 million metric tons Received: May 25, 2023
of textile fiber were produced, with polyester (accounting for Revised: July 6, 2023
52% of the volume share) and cotton (representing 24% of the
volume share) being the two most prevalent fiber types.7−9
However, less than 20% of textile waste (equivalent to 65−92
million tons) is recycled annually,10 with the remaining 80%

© XXXX American Chemical Society https://doi.org/10.1021/acssuschemeng.3c03114


A ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 1. Comparison of approaches to PET and blended fabrics chemical recycling. (A) Challenges in chemical recycling of PET and blended
fabric. (B) Catalytic hydrolysis to access terephthalic acid (TPA) and ethylene glycol (EG) (C) NaOH-promoted hydrolysis of PET/cotton
blended fabrics. (D) Ammonia and CO2-catalyzed glycolysis of PET to BHET and recovery of cotton from PET/blended fabrics. (E) Chemical
structures of PET and cotton and cellulose-based materials and their multiple reactive sites.

polymers.26 Traditionally, Lewis acidic metals have served as The challenge in chemical recycling of blended fabrics lies in
catalysts at high reaction temperatures (160−220 °C),22,27,28 the selective recovery of either fiber type, given their blended
to overcome the substantial reaction barrier and enhance the and intertwined nature.43,44 Acidic media and Lewis acid
solubilization of PET substrates. They often necessitate catalysts can cause the degradation of cotton due to the labile
laborious purification steps to ensure a high purity of BHET, glycosidic bonds (Figure 1E), resulting in a more challenging
minimizing the environmental risk of metal contamina- purification of monomers,45 especially for terephthalic acid
tions.28−30 For instance, Zn(II) (including ZnO) is a (TPA).46 Basic media, such as overstoichiometric amounts of
commonly used catalyst, but it exhibits significant cytotoxicity inorganic bases, can be utilized to depolymerize PET/viscose
in vitro with a greater risk compared to cobalt, iron, copper, blended fabric (filament, PET 30%) into TPA and viscose
through PET hydrolysis (Figure 1C), requiring an additional
and manganese.31−33 One potential solution to prevent metal
reprotonation step.47,48 In this process, mercerization or even
contamination lies in the use of metal-free organocata-
degradation of cotton and cellulose-based materials is
lysts.34−37 Recently, the combination of a strong acid and a essentially unavoidable during the depolymerization reaction.49
strong base (1,5,7-triazabicyclo[4.4.0]dec-5-ene, TBD) has Recently, Zn(OAc)2 has been investigated for the selective
been reported for PET glycolysis.38,39 Ionic liquids have also glycolysis of polyester/cotton blends (at a 5 g scale).50 The
been explored as reaction media in PET glycolysis.40−42 For scalability raises questions including the heavy-metal contam-
these approaches, the price and recycling of the catalysts are ination, considering (1) the additional cost associated with the
still major obstacles to practical application. Thus, there is a use of betaine-based deep eutectic solvents, and (2) the known
need for cheaper, more accessible, and practical catalytic degradation of cotton under glycolysis conditions, such as
protocols to handle the mountainous PET wastes, especially elevated temperatures and the presence of Zn(II) catalysts.51
the PET blended fabrics. Thus, it is crucial to develop a mild catalytic chemical recycling
B https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 2. (A) Conversion of PET to BHET with various bases in catalytic glycolysis with additional 25 mol % of CO2 (red squares and green
square for NH4HCO3) and with only bases under an N2 atmosphere (gray bars). (B) Illustration of positive CO2 effect in glycolysis. Abbreviations:
ethylene glycol (EG), dimethylaminopyridine (DMAP), ethylenediamine (EDA), monoethanolamine (MEA), triethylamine (TEA), 1,5,7-
triazabicyclo[4.4.0]dec-5-ene (TBD), and 1,1,3,3-tetramethylguanidine (TMG).

method for PET/cotton blends under practical conditions, basic ammonia, and water at temperatures above 36 °C.56,57
avoiding the acid- and base-mediated hydrolysis of cotton Our reaction operated under practical, mild conditions (pH ca.
materials.52,53 7−8) and effectively produced high-purity BHET while
We previously demonstrated the successful utilization of recovering cotton fabrics without significant mercerization or
CO2 in catalysis and (bio)polymer degradation. Specifically, hydrolysis. This approach paves the way for large-scale
CO2 can accelerate chemical recycling of Nylon-6,6 through a chemical recycling and the upcycling of blended composites
CO2-catalyzed transamidation reaction and amide bond and PET-containing waste.
activation.54 Upon completion of the reaction, CO2 can be
easily removed without leaving any residue, resulting in a
purification-free process. Building on these findings, our
■ RESULTS AND DISCUSSION
First we attempted to study CO2-mediated glycolysis with PET
objective was to develop a practical, efficient, and scalable bottle substrate, without any prior chemical or thermal
method for recycling PET using catalytic glycolysis with a pretreatment. An application of CO2 in transesterification at
traceless catalyst, CO2, thereby streamlining the green high pressures (147 bar of CO2, at 100−110 °C; 5 h) has been
chemical recycling process. reported by Otsuji et al., which employed ethyl formate,
Herein, we showcase catalytic chemical recycling of blended HCO2Et, and n-octanol, rendering up to a 27% yield of a
materials,55 addressing the challenges associated with glycolysis transesterification product.58 Building on this work, we
of PET in the presence of cotton-based materials. These hypothesized that the addition of a base could enhance
materials present several unresolved challenges, particularly the general base catalysis, while a catalytic quantity of electrophilic
need for selective glycolysis of PET while preserving the cotton (Lewis acidic) CO2 would provide cooperative general acid
fabrics in the blends (Figure 1A,E). Sustainable recovery catalysis under thermal conditions.
methods for cotton are crucial due to its significant carbon An examination of various organic and inorganic bases
footprint, high water usage, and labor issues during production. revealed that the conversion of PET to BHET is related to the
Consequently, there is growing interest in environmentally pKa values of the base catalysts. The maximum conversion was
responsible alternatives for cotton or chemical recycling achieved when employing weak bases (pKa ≈ 6, Figure 2A) in
methods for polyester-blended fabrics (Figure 1D). In this the presence of 25 mol % of CO2. Stronger bases displayed no
study, we demonstrated a traceless and selective catalytic significant improvement with or without CO2. Interestingly,
transesterification of PET/cotton blended fabrics using weak bases such as ammonia and pyridine demonstrated
NH4HCO3 as an inexpensive and easy-to-handle precatalyst exceptional performance, exhibiting enhanced PET degrada-
(up to 25 mol %). This precatalyst decomposed into CO2, tion to BHET in the presence of CO2 (Figure 2B). These
C https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Table 1. PET Bottle Chemical Recyclinga

entry PET (g) catalyst (mol %) conversion (%) yield of BHET (%)
1b 0.2 CO2 (15−30 mol % or 1 atm) not detected 0
2 0.2 N2 not detected 0
3 2 Zn(OAc)2 (25 mol %) 44 40
4 2 ZnCl2 (25 mol %) 75 52
5 50 NH4HCO3 (25 mol %) 88 64 (32 g)
a
Reaction conditions: a PET bottle substrate (0.2−50 g) was used, a catalyst and ethylene glycol were added with a magnetic stirring bar, and the
vial was heated for 6 h. Conversion (%) = (Woligomers+BHET/WPET bottles) × 100; yield of BHET (%) = (WBHET/WPET bottles) × 100; W = weight. The
reaction mixture was diluted with water to precipitate oligomers and BHET, which was crystallized to determine the yield of isolated BHET.
b
Reaction time extended to 24 h.

results suggest that the glycolysis reaction mechanism follows a first demonstration of the synergistic effect of catalytic CO2
general base catalysis, highlighting the crucial role of catalytic and an amine in glycolysis. Solely using ammonia as a catalyst
CO2 in reactions involving weak bases, especially ammonia, without CO2 (aqueous ammonia and an NH3 solution in
generated from NH4HCO3 through thermal decomposition. methanol; Table S1, entry 22) was not effective, yielding
Through further optimization, we successfully established inferior results compared with the NH4HCO3 catalyst,
practical reaction conditions, even in scales ranging from 0.2 to highlighting the importance of CO2. The use of ammonium
50 g, that consistently produced high-purity BHET without the carbonate, (NH4)2CO3, a reagent commonly employed in the
need for complex purification steps (Table 1). Under a N2 Bucherer−Berg reaction,61 yielded results comparable to those
atmosphere and in the absence of catalysts, no background of the ammonium bicarbonate catalyst (87% yield of BHET;
glycolysis was observed (entry 2; see pictures of time- Table S1, entry 18). In contrast, inorganic bases led to the
dependent PET flake degradation with and without a catalyst). contamination of BHET with the hydrolysis product,
The utilization of CO2 alone, whether in catalytic amounts terephthalic acid (TPA), as summarized in Table S1.62
(15−30 mol %) or as a reaction atmosphere (1 atm), had no With our optimized conditions for PET glycolysis, we aimed
discernible effect on the reaction, even with extended reaction to depolymerize PET in the presence of cellulose, as in
times (up to 24 h at 150−200 °C; entry 1). blended fabrics, to separate BHET and recycled cellulose
With this observation, we swiftly conducted comparative (Figure 3). Under acidic conditions, cotton and cellulose-based
studies to evaluate our glycolysis method against existing PET materials are prone to decompose via hydrolysis, as
depolymerization techniques. Zn-based Lewis acidic catalysts, demonstrated in Figure 3A. The catalytic reaction with Zn(II)
such as Zn(OAc)2 and ZnCl2, have been widely used in catalyst resulted in complete decomposition of the cotton
industrial glycolysis of PET and have demonstrated satisfactory fabric within 2 h, with all visible cotton threads disintegrating
conversion of PET to BHET�44% and 75%, respectively after 3 h of reaction time. FT-IR (Fourier-transform infrared
(entries 3 and 4).27,59 However, an elemental analysis of the spectroscopy) showed full conversion of ester functional
isolated BHET from these experiments with Zn2+ revealed groups on PET fabric (Figure 3B; see full spectra in Figure
varying degrees of metal impurities after hot filtration and S18) while the cotton fabric remained intact even after
crystallization of BHET (Table S9), likely due to residual zinc prolonged reaction times under catalytic conditions with
ions. These impurities might hinder their application in the NH4HCO3, (18 h). Solid-state 13C NMR spectra were
repolymerization of BHET to PET.28 obtained to confirm the effective catalytic glycolysis of PET
Under our optimized glycolysis conditions, the use of in the presence of cotton materials (Figure 3C). The spectra
NH4HCO3 as a catalyst allowed us to conduct a larger-scale clearly showed the full conversion of PET and preservation of
reaction (260 mmol, 50 g of waste PET bottle substrate). This the recycled cotton (green), with no presence of aromatic and
catalytic system achieved quantitative conversion of PET, sp2 carbons from PET starting material (black).63
yielding BHET with an isolated yield of 64% in a single The quality of the recovered viscose was further assessed
crystallization step without any byproducts (Table 1, entry using powder X-ray diffraction (PXRD; Figure 3D). The peaks
5).60 This result represents, to the best of our knowledge, the centered at 2θ angles of 12.5, 20, and 22° correspond to the
D https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 3. Catalytic depolymerization of blended textiles and characterization of recovered textile. (A) Catalytic glycolysis of blended fabric and
gravimetric analysis of textile residue. (B) FT-IR analysis of the recovered cotton from the reaction with NH4HCO3 (green), without catalysts
(gray), commercial cotton (orange), and starting material (black). (C) Solid-state 13C NMR spectra and (D) PXRD (powder X-ray diffraction)
patterns (smoothed) of the substrate (orange, viscose) and recovered viscose from glycolysis conditions with ZnCl2 (black, 6 h), NaOH (gray, 19
h), and NH4HCO3 catalysts (green, 19 h). (E) Optical images of the recovered cotton over time, starting from the blended fabrics (47% PET+53%
cotton), with ZnCl2 and NH4HCO3 as catalysts.

crystalline structure of cellulose II. Notably, the PXRD pattern with Zn-based catalysts resulted in complete degradation, with
of the recovered cotton (green) using NH4HCO3 exhibited all fabric threads being degraded within 3 h of reaction time. In
mild crystallization as reported in the literature under basic contrast, NH4HCO3 effectively converted PET to BHET while
conditions with NaOH (gray),49 while the use of ZnCl2 preserving the integrity of the textile thread. Additionally, the
catalyst (black) significantly increased the crystallinity of
analysis using dynamic light scattering (DLS) (Figure S23)
viscose after 6 h.
Based on optical microscopy of the recovered materials and thermogravimetric analysis (Figure S25) confirmed the
(20−80 fold magnification, Figure 3E; also see Figures S20 and same conclusion, further validating the quality of the recovered
S21), it was evident that glycolysis reactions of blended fabrics viscose materials under our depolymerization reaction
E https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 4. Impact of catalysis on the selective chemical recycling of textiles. (A) Catalytic glycolysis of PET containing textile (PET 75%) with and
without the catalyst and the recovery of cotton. (B) Large-scale PET/cellulose-blended textiles for catalytic glycolysis in large scales. (C) Combined
mimicked plastic waste, catalytic glycolysis of PET, and isolation of BHET, cotton, and other plastic waste: polyethylene (PE), polypropylene (PP),
polyurethane (PU), polystyrene (PS), polyoxomethylene (POM), polyvinyl chloride (PVC), polyacrylamide (PAA), acrylonitrile butadiene styrene
(ABS), acrylonitrile butadiene rubber (NBR), stryene-ethylene-butylene-styrene (SEBS), and human hair (peptides).

conditions compared to those obtained using ZnCl2 and our delight, reasonable yields of BHET and quantitative yields
NaOH. for recovered cotton and other fabrics were achieved through a
Considering the diverse range of blended fabric waste to be straightforward separation process (Figure 4B). The purity of
recycled, including various PET:cotton blending ratios, BHET obtained from our catalysis was confirmed by PXRD
unidentifiable dyes,64 additives, and impurities, we subjected and high-performance liquid chromatography (HPLC), with
different commercial textile materials to our catalytic glycolysis no indication of metal impurities or oligomers (Figures S26
without any pretreatment. Reactions performed without the and S27).
catalyst were visibly slower, with textile starting materials still The quality of the isolated BHET was further supported by
remaining in the reaction vessels even after 24 h of reaction
catalyst-free thermal repolymerization to PET (rPET), as
time. In contrast, catalytic glycolysis with NH4HCO3 enabled
analyzed by FT-IR, thermal gravimetric analysis (TGA),
clean conversion of the starting materials to the desired BHET,
even in large-scale reactions without pressure buildup (Figure differential scanning calorimetry (DSC) (Figures S28 and
4A,B). For instance, end-of-life sofa covers consisting of PET/ S29), and elemental analysis (Table S10). Thermal gravimetric
cotton blended fabrics (213 g) yielded 62% and 94% of BHET analysis (Figure S19) and elemental analysis (Table S5) were
and cotton, respectively, without requiring substrate pretreat- also conducted on the residual cellulose and solid materials
ment. from the glycolysis process to corroborate these results.
Furthermore, we evaluated the tolerability of our Although the low whiteness of the obtained BEHT can
NH4HCO3-catalyzed glycolysis method on different blended deteriorate the quality of rPET, a new sublimation method can
fabrics (47−100% PET with cotton, viscose, and lyocell). To effectively improve the whiteness.64
F https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 5. (A) Crystallinity of remaining PET bottle flakes from reaction mixtures under glycolysis conditions (180 °C). (B) Kinetic analysis (170−
190 °C) and activation barriers of the polyester fabric (PET 100%) degradation (eu: cal/(K mol)).

Scheme 1. Control Experiments

To demonstrate the feasibility and robustness of our formation of the precatalyst NH4HCO3 during the reaction
methodology, we mixed several types of polymers (PVC, PP, was confirmed, and the catalyst was easily recovered through
PE, polyamide, PU, PS, ABS, POM, NBR, and SEBS) and even simple sublimation.
included human hair, simulating realistic plastic waste from a To gain insight into the reaction mechanism of NH4HCO3-
waste collection point (Figure 4C). Under our standard mediated PET degradation, we conducted kinetic experiments
reaction conditions, selective glycolysis of PET was achieved, on the glycolysis of PET bottle substrates. The aim was to
with a 23% isolated yield of BHET after a single confirm the catalytic performance of NH4HCO3 compared to
recrystallization step and 75% yield of recovered cotton. The noncatalytic conditions by using thermal gravimetric analysis
other plastic materials were recovered with some deformation (Kissinger model). Our analysis revealed degradation kinetics
but without any signs of chemical depolymerization. The with the NH4HCO3 catalyst comparable to that of the
G https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

industrially widely used ZnCl2 catalytic system for the presumably originating from ammonia. It is plausible that
depolymerization of PET under identical reaction conditions. ammonia can act as a nucleophile in the reaction mixture, with
This finding was surprising considering that ammonia and CO2 the amide intermediate subsequently being converted to
exhibit catalytic activities comparable to strongly Lewis acidic BHET. To investigate this, we subjected monoamide (3)
zinc(II). Additionally, differential scanning calorimetry (DSC) and diamide (5) to the standard conditions (Scheme 1B,C).
analysis of the residual PET flakes collected at different Monotransesterification predominantly produced product 4
reaction times indicated that the PET depolymerization with (75% 1H NMR yield), while the primary amide functional
NH4HCO3 resulted in lower crystallinity of the PET substrates group remained stable, and BHET was obtained as a minor
(Figure 5A, green), while Zn catalysis significantly increased product (25%). Under standard glycolysis conditions for PET,
the crystallinity of the remaining PET substrate under thermal terephthalamide 5 and TPA were converted to BHET in 11%
conditions over time (Figure 5A, gray). These observations and 37% yields, respectively. Consequently, we have ruled out
suggest that the NH4HCO3-catalyzed glycolysis may have reaction pathways involving amides and carboxylic acid
distinct mechanistic features compared with the Zn-catalyzed intermediates resulting from ammonia acting as a nucleophile
process, leading to differences in the crystallinity of the residual (see also DFT calculations, Figure S31). After eliminating a
PET flakes. nucleophilic catalysis mechanism, we propose that ammonia
Therefore, further kinetic studies were performed in terms of and CO2 function as general base and acid catalysts,66−68 as
degradation of textile substrates (100% PET) and their indicated in the experimental data (Figure 2 and see Table S11
conversion into BHET (Figure 5B) at different temperatures. for DFT calculations of various reaction pathways).
For the catalyst ZnCl2, an activation barrier of 20.2 kcal/mol These control experiments implied that CO2 can suppress
was obtained, which is consistent with previously reported the aminolysis of PET. To further verify this, we employed
cases.65 In contrast, our catalyst NH4HCO3 exhibited a higher overstoichiometric amounts of nucleophilic pyrrolidine under
activation barrier (34.9 kcal/mol), as supported by our DFT similar glycolysis conditions (Scheme 1D). The diamide 6 was
(density functional theory) calculations using methyl benzoate a major product (33%) under a catalyst-free N2 atmosphere
as a substrate (see Figure S30). It is noteworthy that a positive due to the high nucleophilicity of pyrrolidine. In contrast,
activation entropy (+3.9 eu (cal/K mol)) was calculated based BHET was detected as the primary product in the presence of
on the Eyring equation for NH4HCO3, contrasting with the pyrrolidine when the reaction was carried out under ambient
results obtained with the Zn(II) catalytic system (−27 eu). CO2 conditions (1 atm). This result unambiguously confirms
This positive entropy value is reasonable for addition/ our hypothesis: CO2-assisted ammonia-catalyzed transester-
elimination reactions, where the addition step can be rate- ification takes place, with CO2 effectively preventing aminolysis
determining, and unfavorable negative activation entropy is and maintaining an optimal concentration of catalytic
commonly observed. In terms of the reaction mechanism, we ammonia in the solution.
postulate that NH4HCO3, as a thermally labile precatalyst,
undergoes decomposition, resulting in an enthalpy loss of 8.2
kcal/mol but a significant entropic gain of 283 cal/(K mol)
(ΔGrxn = −11.9 kcal/mol) due to the liberation of ammonia,
■ CONCLUSION
We have discovered the remarkable advantages of NH4HCO3-
water, and CO2. This entropic gain likely impacts the energy catalyzed selective depolymerization of PET, which allows for
profile of the transesterification reaction, lowering the ΔG⧧ the production of high-purity BHET monomer on a large scale
value. (213 g) while preserving the integrity of cotton and cellulose-
It is important to note that the decomposition of NH4HCO3 based materials. Comprehensive structural analysis of the
and the catalysis are separate events, and further mechanistic recycled cellulose fabrics was performed by using solid-state
13
studies are required to fully understand the relationship C NMR, FT-IR, TGA, DSC, elemental analysis, and
between the entropy gain and the reduction of the free energy microscopy, confirming the quality of the recovered materials.
of activation. Nevertheless, our experimental observations Significant contributions from CO2 as a molecular catalyst,
suggest continuous regeneration of NH4HCO3 via decom- combined with ammonia in the general base-catalyzed PET
position and precipitation on a reflux condenser as ammonium glycolysis, were observed: (1) NH4HCO3, being a more user-
(bi)carbonate, which falls back into the reaction mixture. friendly catalyst precursor due to CO2, eliminates the use of
Qualitative analysis of CO2/ammonia (Figure S21) and toxic and volatile ammonia gas; (2) the presence of CO2
control experiments (Figure S21g) were conducted to confirm effectively suppresses aminolysis reactions; (3) the avoidance
the fate of CO2 and ammonia after completion of the of catalyst contamination in BHET simplifies repolymerization
glycolysis. We detected higher concentrations of ammonia in processes.69
the presence of CO2, confirming our hypothesis on the role of As a result, our approach enables more environmentally
CO2 suppressing ammonia evaporation, in turn effectively friendly chemical processes utilizing NH4HCO3 or other
increasing the catalyst concentration. combinations of organic and inorganic bases with CO2. The
To further support our mechanistic hypothesis regarding synergistic effect of CO2 as a catalyst will be further explored in
CO2 and ammonia-mediated transesterification, we conducted renewable transesterification,70 catalytic polymerization, and
control experiments with the PET bottle substrate and model depolymerization of other plastics, building on the insights
compound 1 (Scheme 1). Trace amounts of amide byproducts gained from our detailed mechanistic studies of PET
(3) were isolated as expected under standard depolymerization depolymerization and cotton recovery. By offering a practical
conditions at a low temperature of 80 °C (Scheme 1A). We and sustainable solution, our method contributes to achieving
found that transesterification with CO2/ammonia was effective the closed-loop production of PET and PET blended fabrics.
even at low temperatures: 57% of BHET was isolated at 80 °C, The utilization of recycled cellulose-based materials and cotton
along with a monofunctionalized product (2, 36%). Interest- also addresses the challenges posed by the fast fashion industry
ingly, amide 3 (<1%) was isolated as an aminolysis byproduct and plastic-based textile waste.8,11,71,72
H https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

■ EXPERIMENTAL SECTION
General Information. All chemicals, unless stated otherwise,
the repeating unit of PET), ethylene glycol (25 equiv), and
ammonium bicarbonate (25 mol %). The reaction mixture was
stirred and heated to 150−200 °C. The conversions (shown in Tables
were purchased from commercial suppliers and used without further
S1 and S2) were detected by analyzing aliquot samples by 1H NMR
purification. CO2 of 99.7% purity was directly used from a CO2
cylinder without any treatment for reactions. Solvents used were of spectroscopy.
high-performance liquid chromatography (HPLC) grade. Ethylene Large-Scale Glycolysis of PET/Cellulose Blended Fabrics. In
glycol was obtained from Sigma-Aldrich. The model compound a three necked round-bottom flask equipped with a reflux condenser
diethyl terephthalate was obtained from TCI. Textile wastes were were added fabrics (47− 100% PET, 80−213 g), ethylene glycol (25
used without any further treatment for the depolymerization equiv, referenced to PET), and NH4HCO3 (25 mol %). Then the
reactions. Drinking bottles were collected from end users. Analytical reaction mixture was heated to 187−195 °C for 18−24 h. Then hot
thin-layer chromatography was done on Merck DC-Alufolien SiO2 60 filtration of the reaction mixture was performed. The filtrate was
F254 0.2 mm thick precoated TLC plates. Column chromatography cooled to room temperature and further to 5 °C overnight. The
was performed using SiO2 from ROCC (SI 1721, 60 Å, 40−63 μm). cotton residual was washed with water, EtOH, and Et2O, respectively,
Liquid 1H and 13C NMR (nuclear magnetic resonance) spectra were and then dried overnight in open air. The BHET crystals were
recorded with a 500 MHz Ultrashield Plus 500 spectrometer and a separated from the mother liquor by filtration as the first crop of
Bruker 125 MHz spectrometer. 1H and 13C NMR spectra were also product. Then the second crop of product was obtained via
recorded at 500 and 126 MHz, respectively, on a Bruker Avance III crystallization with the mother liquor by adding a few crystal seeds
spectrometer with a BBFO probe. All chemical shifts (δ) are quoted and water (water/ethylene glycol = 1:1, v/v).
in ppm, and all coupling constants (J) are expressed in hertz (Hz).
The following abbreviations are used for multiplicity for NMR
resonances: s = singlet, bs = broad singlet, d = doublet, t = triplet, q =
quartet, and m = multiplet. LC-MS (liquid chromatography−mass

*
ASSOCIATED CONTENT
sı Supporting Information
spectroscopy) analyses were carried out by connecting the The Supporting Information is available free of charge at
aforementioned HPLC (high-performance liquid chromatography)
apparatus to a Bruker MicrOTOF-QII system equipped with an ESI
https://pubs.acs.org/doi/10.1021/acssuschemeng.3c03114.
source with nebulizer gas at 1.2 bar, dry gas at 10 L/min, dry Materials and methods, supplementary text, reaction
temperature at 200 °C, capillary at 4500 V, and end plate offset at optimization, characterization, kinetics, NMR spectra,
−500 V. The ion transfer was conducted with funnel 1 and funnel RFs
at 200.0 Vpp and hexapole RF at 100.0 Vpp while the quadrupole ion TGA traces, HPLC traces, PXRD patterns, DSC
energy was set at 5.0 eV with a low mass cutoff at 100.00 m/z. In the analysis, FT-IR spectra, and DFT calculations (PDF)
collision cell, collision energy was set at 8.0 eV and collision RF at
100.0 Vpp, and a transfer time of 80.0 μs and prepulse storage of 1.0 μs
were used. FT-infrared spectra were recorded on a Bruker ALPHA-P
FT-IR spectrometer with a single-reflection ATR module. The
■ AUTHOR INFORMATION
Corresponding Authors
thermogravimetric analysis (TGA) was performed using a Discovery
TGA apparatus from TA Instruments (New Castle, DE, USA) under Gianluca Ciancaleoni − Dipartimento di Chimica e Chimica
a constant flow of 25 mL/min nitrogen. The samples were heated in a Industriale, Università di Pisa, I-56124 Pisa, Italy; CIRCC, I-
platinum TGA pan from room temperature to 600 °C at a heating 70126 Bari, Italy; orcid.org/0000-0001-5113-2351;
rate of 10 °C/min. From the TGA results, the thermal Email: gianluca.ciancaleoni@unipi.it
decompositions and the first derivative of TGA (dTGA) temperatures Ji-Woong Lee − Department of Chemistry and Nanoscience
of the samples were determined using Trios v5.1.1.46572 software Center, University of Copenhagen, 2100 Copenhagen Ø,
(TA Instruments, New Castle, DE, USA). Differential scanning Denmark; Novo Nordisk Foundation CO2 Research Center,
calorimetry (DSC) analysis was performed using a Discovery DSC Aarhus 8000, Denmark; orcid.org/0000-0001-6177-
instrument (TA Instruments, New Castle, DE, USA). 4−5 mg
samples were weighed in Tzero aluminum pans with a perforated lid.
4569; Email: jiwoong.lee@chem.ku.dk
Analyses of the samples were conducted under a nitrogen flow of 50
Authors
mL/min using a linear heating rate of 10 °C/min. The samples were
heated from 30 to 300 °C. The melting point temperatures (Tm) and Yang Yang − Department of Chemistry, University of
melting enthalpies (ΔHm) were determined using Trios v5.1.1.46572 Copenhagen, 2100 Copenhagen Ø, Denmark; orcid.org/
software (TA Instruments, New Castle, DE, USA). The purity of the 0000-0002-7588-1653
isolated BHET (bis(hydroxyethyl)terephthalate) was measured with Shriaya Sharma − Department of Chemistry, University of
HPLC (column Kromasil 5-AmyCoat (4.6 × 250 mm), isopropanol/ Copenhagen, 2100 Copenhagen Ø, Denmark
n-heptane (60/40, v/v), 0.5 mL/min, 8.9 MPa, detector A 220 nm, Carlo Di Bernardo − Department of Chemistry, University of
detector B 254 nm). The X-ray powder diffraction (PXRD) patterns Copenhagen, 2100 Copenhagen Ø, Denmark
were recorded for 2 h using a Bruker AXS D8 powder diffractometer
Elisa Rossi − Dipartimento di Chimica e Chimica Industriale,
(Bruker, Germany) with λ(Cu Kα) = 1.5406 Å radiation (40 kV, 40
mA) covering a 2θ range from 5 to 50°. The 13C solid-state NMR Università di Pisa, I-56124 Pisa, Italy
measurements were performed with a 600 MHz Avance III HD Rodrigo Lima − Department of Chemistry, University of
spectrometer (v13C = 150.9 MHz) equipped with a 4 mm CP/MAS Copenhagen, 2100 Copenhagen Ø, Denmark; orcid.org/
broad-band probe. The reported spectra were acquired with the 0000-0003-1528-845X
13
C{1H} CP-TOSS pulse sequence employing a ramped contact pulse Fadhil S. Kamounah − Department of Chemistry, University
of 1.5 ms (vRF = 55 kHz), an interscan delay of 5 s, a MAS frequency of Copenhagen, 2100 Copenhagen Ø, Denmark
of 5 kHz, and high-power SPINAL64 1H decoupling (vRF = 100 kHz) Margarita Poderyte − Department of Chemistry, University of
during acquisition. 2000 scans were acquired for each spectrum. Copenhagen, 2100 Copenhagen Ø, Denmark
Chemical shifts were referenced relative to TMS using the downfield Kasper Enemark-Rasmussen − DTU Chemistry, Technical
signal of neat adamantane as a secondary reference (δ = 38.48 ppm). University of Denmark, DK-2800 Kgs Lyngby, Denmark
All measurements were performed at 25 °C.
Glycolysis of PET Bottle Flakes. A reactor equipped with a Complete contact information is available at:
magnet was charged with bottle flakes (calculated with molar mass of https://pubs.acs.org/10.1021/acssuschemeng.3c03114
I https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Notes (15) Mandal, S., Dey, A. In Recycling of Polyethylene Terephthalate


The authors declare the following competing financial Bottles; Thomas, S., Rane, A. V., Kanny, K., Abitha, V. K., Thomas, M.
interest(s): J.-W.L., Y.Y., S.S., C.D.B., and F.S.K. filed a patent G., Eds.; William Andrew Publishing: 2019; pp 1−22.
(16) Kosloski-Oh, S. C.; Wood, Z. A.; Manjarrez, Y.; de los Rios, J.
application (EP 23150108.1). There are no other interests.
P.; Fieser, M. E. Catalytic Methods for Chemical Recycling or

■ ACKNOWLEDGMENTS
This paper is dedicated to the Department of Chemistry,
Upcycling of Commercial Polymers. Mater. Horiz. 2021, 8, 1084−
1129.
(17) Sheel, A.; Pant, D. In Recycling of Polyethylene Terephthalate
Bottles; Thomas, S., Rane, A. V., Kanny, K., Abitha, V. K., Thomas, M.
Sungkyunkwan University, on the occasion of its 70th G., Eds.; William Andrew Publishing: 2019; pp 61−84.
anniversary. The generous support from the Department of (18) Martín, A. J.; Mondelli, C.; Jaydev, S. D.; Pérez-Ramírez, J.
Chemistry, University of Copenhagen, the Carlsberg Founda- Catalytic Processing of Plastic Waste on the Rise. Chem. 2021, 7,
tion (CF21-0308), NNF CO2 Research Center (CORC), and 1487−1533.
the Novo Nordisk Foundation (NNF20OC0064347) is (19) de Dios Caputto, M. D.; Navarro, R.; Valentín, J. L.; Marcos-
gratefully acknowledged. We thank Prof. Troels Skrydstrup Fernández, Á . Chemical Upcycling of Poly(ethylene terephthalate)
for the fruitful discussion on depolymerization and Dr. Theis Waste: Moving to a Circular Model. J. Polym. Sci. 2022, 60, 3269−
Brock-Nannestad, Christian Tortzen, and Prof. Pernille Harris 3283.
for the analysis and acquisition of optical microscope images. (20) Raheem, A. B.; Noor, Z. Z.; Hassan, A.; Abd Hamid, M. K.;
Samsudin, S. A.; Sabeen, A. H. Current Developments in Chemical
The NMR Center • DTU and the Villum Foundation are Recycling of Post-Consumer Polyethylene Terephthalate Wastes for
acknowledged for allowing us access to the 600 MHz New Materials Production: A review. J. Clean. Prod. 2019, 225, 1052−
spectrometer. We acknowledge Prof. Heloisa Nunes Bordallo 1064.
and CPHarma, Department of Pharmacy, Faculty of Health (21) George, N.; Kurian, T. Recent Developments in the Chemical
and Medical Sciences, University of Copenhagen, for the Recycling of Postconsumer Poly(ethylene terephthalate) Waste. Ind.
access to TGA and DSC measurements. We also thank our Eng. Chem. Res. 2014, 53, 14185−14198.
analytical departments for their kind support. (22) Sinha, V.; Patel, M. R.; Patel, J. V. PET Waste Management by
Chemical Recycling: A Review. J. Polym. Environ. 2010, 18, 8−25.

■ REFERENCES
(1) Geyer, R. In Plastic Waste and Recycling; Letcher, T. M., Ed.;
(23) Maurya, A.; Bhattacharya, A.; Khare, S. K. Enzymatic
Remediation of Polyethylene Terephthalate (PET)−Based Polymers
for Effective Management of Plastic Wastes: An Overview. Front.
Academic Press: 2020; pp 13−32. Bioeng. Biotechnol. 2020, 8, 602325.
(2) Ryberg, M. W.; Hauschild, M. Z.; Wang, F.; Averous-Monnery, (24) Lu, H.; Diaz, D. J.; Czarnecki, N. J.; Zhu, C.; Kim, W.; Shroff,
S.; Laurent, A. Global environmental losses of plastics across their R.; Acosta, D. J.; Alexander, B. R.; Cole, H. O.; Zhang, Y.; Lynd, N.
value chains. Resour., Conserv. Recycl. 2019, 151, 104459. A.; Ellington, A. D.; Alper, H. S. Machine Learning-aided Engineering
(3) Towards a True Circular Economy of PET Plastics and Textiles of Hydrolases for PET Depolymerization. Nature 2022, 604, 662−
thanks to Enzymatic Recycling of Waste, https://webgate.ec.europa.eu/ 667.
life/publicWebsite/project/details/5731#. (25) Yoshida, S.; Hiraga, K.; Takehana, T.; Taniguchi, I.; Yamaji, H.;
(4) Barnard, E.; Rubio Arias, J. J.; Thielemans, W. Chemolytic Maeda, Y.; Toyohara, K.; Miyamoto, K.; Kimura, Y.; Oda, K. A
depolymerisation of PET: a review. Green Chem. 2021, 23, 3765− Bacterium That Degrades and Assimilates Poly (ethylene tereph-
3789. thalate). Science 2016, 351, 1196−1199.
(5) Palakurthi, M. Development of composites from waste PET-cotton (26) Mendiburu-Valor, E.; Calvo-Correas, T.; Martin, L.;
textiles; Master’s Thesis, University of Nebraska, Lincoln, NE, USA, Harismendy, I.; Peña-Rodriguez, C.; Eceiza, A. Synthesis and
December 2016. Characterization of Sustainable Polyurethanes from Renewable and
(6) Palacios-Mateo, C.; van der Meer, Y.; Seide, G. Analysis of the Recycled Feedstocks. J. Clean. Prod. 2023, 400, 136749.
Polyester Clothing Value Chain to Identify Key Intervention Points (27) Delle Chiaie, K. R.; McMahon, F. R.; Williams, E. J.; Price, M.
for Sustainability. Environ. Sci. Eur. 2021, 33, 2. J.; Dove, A. P. Dual-Catalytic Depolymerization of Polyethylene
(7) Wang, S.; Salmon, S. Progress toward Circularity of Polyester Terephthalate (PET). Polym. Chem. 2020, 11, 1450−1453.
and Cotton Textiles. Sustain. Chem. 2022, 3, 376−403. (28) Zhang, Q.; Huang, R.; Yao, H.; Lu, X.; Yan, D.; Xin, J. Removal
(8) Niinimäki, K.; Peters, G.; Dahlbo, H.; Perry, P.; Rissanen, T.; of Zn 2+ from Polyethylene Terephthalate (PET) Glycolytic
Gwilt, A. The Environmental Price of Fast Fashion. Nat. Rev. Earth Monomers by Sulfonic Acid Cation Exchange Resin. J. Environ.
Environ. 2020, 1, 189−200. Chem. Eng. 2021, 9, 105326.
(9) Guo, Z.; Eriksson, M.; Motte, H. d. l.; Adolfsson, E. Circular (29) Benyathiar, P.; Kumar, P.; Carpenter, G.; Brace, J.; Mishra, D.
Recycling of Polyester Textile Waste using a Sustainable Catalyst. J. K. Polyethylene Terephthalate (PET) Bottle-to-Bottle Recycling for
Clean. Prod. 2021, 283, 124579. the Beverage Industry: A Review. Polymers 2022, 14, 2366−2394.
(10) Egan, J.; Salmon, S. Strategies and Progress in Synthetic Textile (30) Giannotta, G.; Po, R.; Cardi, N.; Tampellini, E.; Occhiello, E.;
Fiber Biodegradability. SN Appl. Sci. 2021, 4, 1−36. Garbassi, F.; Nicolais, L. Processing Effects on Poly(ethylene
(11) A new Textile Economy: Redesigning Fashion’s future; Banks, I., terephthalate) from Bottle Scraps. Polym. Eng. Sci. 1994, 34, 1219−
Gravis, L., Eds; Ellen MacArthur Foundations: 2017. 1223.
(12) Meys, R.; Frick, F.; Westhues, S.; Sternberg, A.; Klankermayer, (31) Borovanský, J.; Riley, P. A. Cytotoxicity of Zinc in vitro. Chem.
J.; Bardow, A. Towards a Circular Economy for Plastic Packaging Biol. Interact. 1989, 69, 279−291.
Wastes − the Environmental Potential of Chemical Recycling. Resour. (32) Ghanmi, Z.; Rouabhia, M.; Othmane, O.; Deschaux, P. A.
Conserv. Recycl. 2020, 162, 105010. Effects of metal ions on cyprinid fish immune response: In vitro
(13) Chu, M.; Liu, Y.; Lou, X.; Zhang, Q.; Chen, J. Rational Design Effects of Zn2+ and Mn2+ on the Mitogenic Response of Carp
of Chemical Catalysis for Plastic Recycling. ACS Catal. 2022, 12, Pronephros lymphocytes. Ecotoxicol. Environ. Saf. 1989, 17, 183−189.
4659−4679. (33) Bondarenko, O.; Juganson, K.; Ivask, A.; Kasemets, K.;
(14) Zhang, F.; Zeng, M.; Yappert, R. D.; Sun, J.; Lee, Y.-H.; Mortimer, M.; Kahru, A. Toxicity of Ag, CuO and ZnO Nanoparticles
LaPointe, A. M.; Peters, B.; Abu-Omar, M. M.; Scott, S. L. to Selected Environmentally Relevant Test Organisms and Mamma-
Polyethylene Upcycling to long-chain Alkylaromatics by Tandem lian Cells in vitro: a Critical Review. Arch. Toxicol. 2013, 87, 1181−
Hydrogenolysis/aromatization. Science 2020, 370, 437−441. 1200.

J https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

(34) Jehanno, C.; Pérez-Madrigal, M. M.; Demarteau, J.; Sardon, H.; (52) Asaadi, S.; Hummel, M.; Hellsten, S.; Härkäsalmi, T.; Ma, Y.;
Dove, A. P. Organocatalysis for Depolymerisation. Polym. Chem. Michud, A.; Sixta, H. Renewable High-Performance Fibers from the
2019, 10, 172−186. Chemical Recycling of Cotton Waste Utilizing an Ionic Liquid.
(35) Fukushima, K.; Coady, D. J.; Jones, G. O.; Almegren, H. A.; ChemSusChem 2016, 9, 3250−3258.
Alabdulrahman, A. M.; Alsewailem, F. D.; Horn, H. W.; Rice, J. E.; (53) Juanga-Labayen, J. P.; Labayen, I. V.; Yuan, Q. A Review on
Hedrick, J. L. Unexpected Efficiency of Cyclic Amidine Catalysts in Textile Recycling Practices and Challenges. Textiles 2022, 2, 174−
Depolymerizing Poly(ethylene terephthalate). J. Polym. Sci., Part A: 188.
Polym. Chem. 2013, 51, 1606−1611. (54) Yang, Y.; Liu, J.; Kamounah, F. S.; Ciancaleoni, G.; Lee, J.-W. A
(36) Fukushima, K.; Coulembier, O.; Lecuyer, J. M.; Almegren, H. CO2-Catalyzed Transamidation Reaction. J. Org. Chem. 2021, 86,
A.; Alabdulrahman, A. M.; Alsewailem, F. D.; Mcneil, M. A.; Dubois, 16867−16881.
P.; Waymouth, R. M.; Horn, H. W.; Rice, J. E.; Hedrick, J. L. (55) Cellulose I, such as native cotton, could be converted to
Organocatalytic Depolymerization of Poly(ethylene terephthalate). J. cellulose II (discussed in this manuscript) via mercerization process,
Polym. Sci. A Polym. Chem. 2011, 49, 1273−1281. see; Kalaoglu, F.; Paul, R. 14 - Finishing of jeans and quality control.
(37) Fan, C.; Zhang, L.; Zhu, C.; Cao, J.; Xu, Y.; Sun, P.; Zeng, G.; In Denim; Paul, R., Ed.; Woodhead Publishing, 2015; pp 425−459.
Jiang, W.; Zhang, Q. Efficient Glycolysis of PET Catalyzed by a (56) van Schijndel, J.; Canalle, L. A.; Molendijk, D.; Meuldijk, J.
Metal-free Phosphazene Base: the Important Role of EG. Green Chem. Exploration of the Role of Double Schiff Bases as Catalytic
2022, 24, 1294−1301. Intermediates in the Knoevenagel Reaction of Furanic Aldehydes:
(38) Jehanno, C.; Flores, I.; Dove, A. P.; Müller, A. J.; Ruipérez, F.; Mechanistic Considerations. Synlett 2018, 29, 1983−1988.
Sardon, H. Organocatalysed Depolymerisation of PET in a Fully (57) van Schijndel, J.; Canalle, L. A.; Molendijk, D.; Meuldijk, J. The
Green Knoevenagel Condensation: Solvent-free Condensation of
Sustainable Cycle using Thermally Stable Protic Ionic Salt. Green
Benzaldehydes. Green Chem. Lett. Rev. 2017, 10, 404−411.
Chem. 2018, 20, 1205−1212.
(58) Otsuji, Y.; Matsumura, N.; Imoto, E. The Carbon Dioxide-
(39) Jehanno, C.; Demarteau, J.; Mantione, D.; Arno, M. C.;
Catalyzed Ester Exchange Reaction. Bull. Chem. Soc. Jpn. 1971, 44,
Ruipérez, F.; Hedrick, J. L.; Dove, A. P.; Sardon, H. Selective
852−854.
Chemical Upcycling of Mixed Plastics Guided by a Thermally Stable (59) Wang, Y.; Zhang, Y.; Song, H.; Wang, Y.; Deng, T.; Hou, X.
Organocatalyst. Angew. Chem., Int. Ed. 2021, 60, 6710−6717. Zinc-Catalyzed Ester Bond Cleavage: Chemical Degradation of
(40) Wang, H.; Li, Z.; Liu, Y.; Zhang, X.; Zhang, S. Degradation of Polyethylene Terephthalate. J. Clean. Prod. 2019, 208, 1469−1475.
Poly(ethylene terephthalate) Using Ionic Liquids. Green Chem. 2009, (60) It should be noted here that, under the tested catalytic
11, 1568−1575. conditions, there was no formation of cyclic carbonate or urea from
(41) Al-Sabagh, A. M.; Yehia, F. Z.; Eissa, A.-M. M. F.; Moustafa, M. CO2 and ethylene glycol.
E.; Eshaq, G.; Rabie, A.-R. M.; ElMetwally, A. E. Glycolysis of (61) Bucherer, H. T.; Lieb, V. A. Ü ber die Bildung substituierter
Poly(ethylene terephthalate) Catalyzed by the Lewis Base Ionic Hydantoine aus Aldehyden und Ketonen. Synthese von Hydantoinen.
Liquid [Bmim][OAc]. Ind. Eng. Chem. Res. 2014, 53, 18443−18451. J. Prakt. Chem. 1934, 141, 5−43.
(42) Wang, L.; Nelson, G. A.; Toland, J.; Holbrey, J. D. Glycolysis of (62) López-Fonseca, R.; Duque-Ingunza, I.; de Rivas, B.; Flores-
PET Using 1,3-Dimethylimidazolium-2-Carboxylate as an Organo- Giraldo, L.; Gutiérrez-Ortiz, J. I. Kinetics of Catalytic Glycolysis of
catalyst. ACS Sustain. Chem. Eng. 2020, 8, 13362−13368. PET Wastes with Sodium Carbonate. Chem. Eng. J. 2011, 168, 312−
(43) Bouwhuis, G. H.; Brinks, G. J.; Groeneveld, R. A. J.; Oelerich, 320.
J., Separation and Recycling of Cotton from Cotton/PET Blends by (63) Palme, A.; Peterson, A.; de la Motte, H.; Theliander, H.; Brelid,
Depolymerization of PET Catalyzed by Bases and Ionic Liquids. H. Development of an Efficient Route for Combined Recycling of
Saxion University of Applied Sciences: 2014. https://hbo-kennisbank. PET and Cotton from Mixed Fabrics. Text. Cloth. Sustain. 2017, 3, 4.
nl/details/saxion_kenniscentra:9DAABF8B-BDE2-4C7C- (64) During the revision of the manuscript, a study on glycolysis of
A96A05B0524BA595. PET textile showed that the dyes could be easily removed. Chen, Z.;
(44) CHARBONNIER, B. Method for Separating Polyester and Sun, H.; Kong, W.; Chen, L.; Zuo, W. Closed-loop Utilization of
Cotton in order to Recycle Textile Waste. WO2013182801A1. Polyester in the Textile Industry. Green Chem. 2023, 25, 4429.
(45) Johnson, S.; Echeverria, D.; Venditti, R.; Jameel, H.; Yao, Y. (65) Chen, F.; Zhou, Q.; Bu, R.; Yang, F.; Li, W. Kinetics of
Supply Chain of Waste Cotton Recycling and Reuse: A Review. Poly(ethylene terephthalate) Fiber Glycolysis in Ethylene Glycol.
AATCC J. Res. 2020, 7, 19−31. Fibers Polym. 2015, 16, 1213−1219.
(46) Han, M. In Recycling of Polyethylene Terephthalate Bottles; (66) Qin, Y.; Zhang, T.; Ching, H. Y. V.; Raman, G. S.; Das, S.
Thomas, S., Rane, A. V., Kanny, K., Abitha, V. K., Thomas, M. G., Integrated Strategy for the Synthesis of Aromatic Building Blocks via
Eds.; William Andrew Publishing: 2019; p 91. Upcycling of Real-Life Plastic Wastes. Chem. 2022, 8, 2472−2484.
(47) Bengtsson, J.; Peterson, A.; Idström, A.; de la Motte, H.; (67) Sahoo, P. K.; Zhang, Y.; Das, S. CO2-Promoted Reactions: An
Jedvert, K. Chemical Recycling of a Textile Blend from Polyester and Emerging Concept for the Synthesis of Fine Chemicals and
Viscose, Part II: Mechanism and Reactivity during Alkaline Hydrolysis Pharmaceuticals. ACS Catal. 2021, 11, 3414−3442.
of Textile Polyester. Sustainability 2022, 14, 6911. (68) Schilling, W.; Das, S. CO2-catalyzed/promoted Transformation
(48) Peterson, A.; Wallinder, J.; Bengtsson, J.; Idström, A.; Bialik, of Organic Functional Groups. Tetrahedron Lett. 2018, 59, 3821−
M.; Jedvert, K.; de la Motte, H. Chemical Recycling of a Textile Blend 3828.
from Polyester and Viscose, Part I: Process Description, Character- (69) Sheel, A.; Pant, D., in Recycling of Polyethylene Terephthalate
ization, and Utilization of the Recycled Cellulose. Sustainability 2022, Bottles; Thomas, S., Rane, A. V., Kanny, K., Abitha, V. K., Thomas, M.
14, 7272. G., Eds.; William Andrew Publishing: 2019; p 74.
(49) Borysiak, S.; Garbarczyk, J. Applying the WAXS Method to (70) Hoydonckx, H. E.; De Vos, D. E.; Chavan, S. A.; Jacobs, P. A.
Estimate the Supermolecular Structure of Cellulose Fibres After Esterification and Transesterification of Renewable Chemicals. Top.
Mercerisation. Fibres Text. East. Eur. 2003, 11, 104−106. Catal. 2004, 27, 83−96.
(50) Liu, L.; Yao, H.; Zhou, Q.; Yao, X.; Yan, D.; Xu, J.; Lu, X. (71) Douglass, E. F.; Avci, H.; Boy, R.; Rojas, O. J.; Kotek, R. A
Review of Cellulose and Cellulose Blends for Preparation of Bio-
Recycling of Full Components of Polyester/Cotton Blends Catalyzed
derived and Conventional Membranes, Nanostructured Thin Films,
by Betaine-based Deep Eutectic Solvents. J. Environ. Chem. Eng. 2022,
and Composites. Polym. Rev. 2018, 58, 102−163.
10, 107512.
(72) Prado, K. S.; Gonzales, D.; Spinace, M. A. Recycling of Viscose
(51) Amarasekara, A. S.; Ebede, C. C. Zinc Chloride Mediated
Yarn Waste through one-step Extraction of Nanocellulose. Int. J. Biol.
Degradation of Cotton at 200°C and Identification of the Products. Macromol. 2019, 136, 729−737.
Bioresour. Technol. 2009, 100, 5301−5304.

K https://doi.org/10.1021/acssuschemeng.3c03114
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

View publication stats

You might also like