Next Article in Journal
Nitrogen-Doped Pitch-Based Activated Carbon Fibers with Multi-Dimensional Metal Nanoparticle Distribution for the Effective Removal of NO
Next Article in Special Issue
Mono(imidazolin-2-iminato) Hafnium Complexes: Synthesis and Application in the Ring-Opening Polymerization of ε-Caprolactone and rac-Lactide
Previous Article in Journal
Insights of Selective Catalytic Reduction Technology for Nitrogen Oxides Control in Marine Engine Applications
Previous Article in Special Issue
Ring Opening Polymerization of Lactides and Lactones by Multimetallic Titanium Complexes Derived from the Acids Ph2C(X)CO2H (X = OH, NH2)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Contribution of Commercial Metal Amides to the Chemical Recycling of Waste Polyesters

Department of Chemistry and Biology “Adolfo Zambelli”, University of Salerno, Via Giovanni Paolo II, 132, 84084 Fisciano, SA, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1193; https://doi.org/10.3390/catal12101193
Submission received: 7 September 2022 / Revised: 27 September 2022 / Accepted: 6 October 2022 / Published: 8 October 2022
(This article belongs to the Special Issue Catalysts for the Ring Opening Polymerization)

Abstract

:
Simple and commercially available metal amides are investigated as catalysts for the chemical depolymerization of polyesters of commercial interest such as polylactide (PLA) and polyethylene terephthalate (PET) via alcoholysis. In the alcoholysis reactions performed with methanol or ethanol at room temperature, Zn, Mg, and Y amides showed the highest activities, while the amides of group 4 metals were revealed as poor catalysts. During the ethanolysis of PLA at higher temperatures and the glycolysis of PET, the good activity of the Zn amide was preserved, while for Mg and Y amides, a significant decrease was observed. The reaction temperature had an opposite effect on the performance of group 4 amides, with the Zr amide revealed to be the best catalyst in the PET glycolysis, reaching activities comparable to the best ones reported in the literature for metal catalysts (78% BHET yield within one hour at 180 °C). These studies represent new opportunities for the sustainable recycling of plastics, which are currently being used on a large scale, and provide significant contributions to the design of a circular economy model in the plastic industry.

Graphical Abstract

1. Introduction

The huge volumes of plastic that are placed on the market for single-use products every year, in combination with the absence of a correct management strategy for their end-of-life, have caused dramatic problems of environmental pollution, demonstrating the unsustainability of the current linear productive model in the long term [1,2,3]. However, due to their versatility and low costs, plastics are irreplaceable materials, and the hypothesis of a plastic-free society is unrealistic [4,5].
To overcome this dichotomy, the efforts of scientific research have focused on the development of both more sustainable plastics, biodegradable and/or obtainable from renewable resources [6,7,8], and strategies for their chemical recycling [9]. In this context, aliphatic polyesters represent the most promising materials as alternatives to traditional petroleum-based polyolefins [10].
In the last two decades, the ring-opening polymerization (ROP) of lactones has received a great impetus as a method of election for the synthesis of polyesters [7,11]. Simple amides or alkoxides of several metals were tested as catalysts for ROP only in initial investigations, showing activity and control efficiency lower than those obtained with the industrial catalyst Sn (Oct)2 [12,13,14,15,16].
Much more numerous are the studies concerning the catalysts based on coordination complexes in which the central metal is coordinated to an ancillary ligand and to one or more labile ligands that act as the initiating groups because these showed improved abilities in terms of the control of polymerization processes [17,18]. For these systems, the ancillary ligands have a crucial role in influencing the catalytic behavior of reactive centers and promoting punctual control over macromolecular parameters, namely the chain-end group structures, molecular masses and their distributions, and stereoregularity [19,20,21,22]. On the other hand, the synthesis of architectonically sophisticated ligands has a profound impact on the costs and sustainability of the whole catalytic process. Thus, the use of catalysts based on non-toxic metals with simple structures and/or commercially accessible is highly desirable.
However, to design a virtuous life cycle for plastics, in addition to greener processes for their production, adequate treatments of end-of-life plastic waste must be planned [23,24,25,26,27].
Beyond ecological considerations, an estimated economic loss of 120 billion dollars per year emanates from plastic waste, which currently cannot be recycled [28]. While the mechanical recycling of plastic waste provides recycled materials with lower quality, the chemical recycling and upcycling of polymers may offer new opportunities to contain the economic and environmental costs of this sector.
Some recent studies described zinc and magnesium complexes as efficient catalysts for the degradation of polyesters [29,30,31,32,33]. Among these, Wang reported zinc bis[bis(trimethylsilyl)amide], (Zn(HMDS)2), as a simple and efficient ligand-free metal catalyst for the ring-opening polymerization of various lactones [34] and for the degradation of polylactide or mixed commercial plastics to obtain value-added molecules [35].
Inspired by these considerations, we explored the catalytic behavior of other commercially available amides of non-toxic and low-cost metals (Scheme 1) in the degradation reactions of polylactide (PLLA).
Using glycolysis, these studies were also extended to the chemical degradation of polyethylene terephthalate (PET), the most important commercial polyester whose annual global production exceeded 70 million tons, mainly used for the production of single-use products. Due to the lack of an efficient recycling capacity for this volume of consumption, PET recycling has become a priority for the preservation of resources and the environment [36].

2. Results and Discussion

2.1. Degradation of PLA via Alcoholysis

The depolymerization of PLA promoted by low-molecular-weight alcohols, such as methanol or ethanol, proceeds through transesterification reactions in which the metal center behaves as a Lewis acid activating the carbonylic groups of polymer chains for the nucleophilic attack by the alcohol (Scheme 2). In this study, using alcoholysis, we investigated the role of metal amides in the depolymerization of end-of-life PLLA plastic cups to alkyl lactates.
The alcoholysis reactions of commercial PLA plastic cups with a molecular mass of 58 KDa were performed at room temperature without additional solvents to minimize the environmental impact of the procedure (see degradation experiments in Table 1).
The PLA degradations were monitored through the NMR analysis. The conversion of internal methine units (Xint), the selectivity of Me-La (SMe-La), and the yield of Me-La (YMe-La) were calculated by the integration of the diagnostic signals of the 1H NMR spectra (Figure 1).
Among the catalysts investigated, the best performance was achieved with the magnesium amide (run 2, Table 1) for which the almost complete degradation of PLA was achieved after only two hours. Its activity was significantly higher than that of the Zn complex, which, after the same time, reached a conversion of 40% (run 2, Table 1). Moreover, the activity of Mg amide was significantly higher than that obtained with the homoleptic Mg complexes supported by phenoxy–imine amine (NNO) ligands that converted up to 93% of a sample of PLA cup dissolved in THF at 80 °C [31,32]. These results indicate that the Mg compound may be used effectively for catalytic degradation at room temperature, which provides high economic benefits.
With the yttrium amide, complete degradation was achieved only after 19 h (run 3, Table 1). Significantly lower activities were observed with group 4 amides (runs 4–6, Table 1), and the Zr compound showed the worst results.
Generally, PLA degradation occurs via a two-step process in which the polymer undergoes a random scission of polymer chains into oligomeric species that are progressively converted into methyl lactate.
Surprisingly, Zn, Mg, and Y catalysts promoted selective processes in which the polymer chain is directly converted into the final product (the percentage of methyl lactate was the same as the degraded PLA). This behavior was previously observed for other catalysts [29,37] that, in the absence of a solvent, promoted a mechanism in which the degradation occurs via the progressive aggression of the chain ends with the direct formation of methyl lactate.
Subsequently, the same metal compounds were tested in terms of the degradation of PLA with ethanol (Figure 2). The final product of the ethanolysis of PLA is ethyl lactate (EtLa), described as a biodegradable “green” solvent with low toxicity, which can be used as an alternative to acetone or toluene [38].
As expected, the degradation of PLA was definitively slower when performed in neat ethanol (runs 1–6, Table 2) than in methanol because the more sterically hindered alkyl group disfavors the approach toward the carbonyl group [39]. At room temperature, almost quantitative conversions were obtained for the amides of Zn, Mg, and Y after 24 h. Additionally, in this case, lower activities were observed with group 4 metals, especially with Zr and Hf.
At higher temperatures, the reactivity trend was significantly different: The Zn amide preserved its activity, while an important downfall was observed for Mg and Y catalysts (runs 7–12, Table 2). The Mg amide showed a performance significantly poorer than the homoleptic complexes reported by Jones and Wood [39]. This could suggest that the presence of ancillary ligands is important to stabilize the metal center when drastic reaction conditions are applied.
Unexpectedly, at higher temperatures, group 4 metal compounds improved their performances. The additional experiments performed at the intermediate temperature of 60 °C (runs 13 and 14, Table 2) confirmed the opposite effect of the temperature on the catalytic activity of yttrium and zirconium catalysts (Figure 3).
These results showed that the trend in the performances of the explored catalysts strongly depended on reaction conditions, likely because of the different thermal stability of the active species involved in the polymerization process.
To exclude the presence of epimerization side reactions through the deprotonation of the beta-CH group by amides, the ethylactate samples obtained through degradation processes were analyzed using optical rotation measurements in methylene dichloride at 22 °C.
The full coherence between the [α]22D value of the ethylactate sample obtained in run 2 of Table 2 ([α]22D = −0.2169) and that of the commercial product (−0.2919) confirmed the absence of side reactions of epimerization.

2.2. Glycolysis of PET

Following PLA degradation success, we decided to extend our studies to the alcoholysis reaction of other polyesters of high commercial interest such as polyethylene terephthalate (PET). Currently, PET accounts for about 23% of plastic use for short-time packaging; thus, improvement in the strategies for its recycling is of fundamental importance.
The degradation of PET via glycolysis is an efficient procedure to convert PET waste into a mixture of bis(2-hydroxyethyl)terephthalate (BHET) and ethylene glycol (EG), which are the monomeric units for the synthesis of PET via polycondensation (Scheme 3).
A variety of metals [40,41] and organic catalysts [42,43,44] have been reported for the chemical depolymerization of PET. Recently, bicomponent catalysts formed by organic bases with Lewis acidic metal salts have also been explored [45]. Usually, because of its scarce solubility and high stability, the alcoholysis of PET requires severe reaction conditions such as high temperature and pressure and elevated percentage of catalyst.
All the metal amides reported in Scheme 1 were investigated in the degradation of a commercial sample of PET (Mn = 42.000 g mol−1, Ð = 1.7). The reactions were performed in the absence of additional solvents at 180 °C and were stopped when the complete dissolution of solid PET was observed (runs 1–6, Table 3). The conversion of PET was evaluated as a ratio of the weight of the decomposed PET (i.e., the difference between the initial weight of PET and the weight of residual PET) and the weight of the initial sample of PET. The yield and selectivity in the production of BHET were estimated as described in the experimental section.
The degradation products were extracted with distilled water and crystallized from cold water. The purity of the BHET obtained via degradation was evaluated with 1H NMR spectroscopy (Figure 4) and mass spectrometry using either electrospray (ESI) or Matrix Assisted Laser Desorption Ionization—Time of Flight (MALDI-ToF) methods (Figure 5).
In the ESI-MS spectra, three peaks were evident: two were related to [BHET + Na+] m/z = 277.06 and [BHET + H+] m/z = 255.08 ions and the other one corresponding to the [BHET − H2O + H+] ion. (Figure 5)
In the MALDI-ToF spectra, a single peak was evident for the [BHET − H2O + H+] ion m/z = 255.08.
Among the investigated catalysts, Zn, Mg, and Y were able to almost quantitatively convert the whole PET amount after one hour. However, the best results were obtained with Zr amide, which reached a good percentage of degradation (86%) and very high selectivity (78%) in the production of BHET, with values that are comparable with the best results obtained with metal catalysts.
Since, in the presence of a large amount of alcohol, the metal amides are reasonably converted into the related metal alkoxides, a comparison experiment was performed by using Zr(OEt)4 as a catalyst, under the same reaction conditions. In this case, under the same reaction conditions, a much lower percentage of the degradation of PET was achieved (19%).
The very good performances shown by Zn amide, in comparison to the related alkoxide, could be a consequence of the nature of its active species and/or of the co-presence of the free amine produced in situ that could activate the alcohol compound through the H bond interaction, as proposed for bicomponent catalysts formed by metal halides and amines described by Dove [45].
To the best of our knowledge, these studies represent the first examples of a systematic exploration of commercial metal amides in PET glycolysis.

3. Experimental Section

3.1. Materials and Methods

All the manipulations of air- and/or water-sensitive compounds were carried out under a dry nitrogen atmosphere using a Braun LabMaster glovebox or standard Schlenk line techniques. The glassware and vials used in the polymerization were dried in an oven at 120 °C overnight and exposed three times to vacuum–nitrogen cycles.
Methanol and ethanol were refluxed over Na and distilled under nitrogen. Benzene, hexane, and tetrahydrofuran were distilled under nitrogen over sodium benzophenone. (THF).
Deuterated solvents, CDCl3, and C6D6 were purchased from Eurisotop were dried over molecular sieves. All the other reagents and solvents were purchased from Aldrich and used without further purification. The metal amides were purchased from Aldrich and used as received. L-lactide was purchased from Aldrich and crystallized using dry toluene and afterward stored at −20 °C in a glovebox. All the other chemicals were commercially available and used as received unless otherwise stated.

3.2. NMR Analysis

The NMR spectra were recorded on Bruker Advance 300, 400, and 600 MHz spectrometers (1H: 300.13, 400.13, 600.13 MHz) at 25 °C unless otherwise stated. Chemical shifts (δ) are expressed as parts per million and coupling constants (J) in hertz.
The resonances are reported in ppm (δ) and the coupling constants in Hz (J) and were referred to with the residual solvent peaks at δ = 7.16 ppm for C6D6 and δ = 7.27 for CDCl3.
The 13C NMR spectra were referred to using the residual solvent peaks at δ = 128.06 for C6D6 and δ = 77.23 for CDCl3. Spectra recording was performed using Bruker’s TopSpin v2.1 software. Data processing was performed using TopSpin v2.1 or MestReNova v6.0.2 software.

3.3. GPC Size Exclusion Chromatography

The number-average molecular weights (Mn) and molecular weight distributions of polymers (dispersity, Ð) were evaluated through size exclusion chromatography (SEC), using Agilent 1260 Infinity Series GPC (ResiPore 3 μm, 300 × 7.5 mm, 1.0 mL min−1, UV (250 nm) and a refractive index (RI, PLGPC 220) detector. All the measurements were performed with THF as the eluent at a flow rate of 1.0 mL/min at 35 °C. Monodisperse poly(styrene) polymers were used as calibration standards.
MALDI-ToF-MS analysis was performed on a Waters MALDI Micro MX equipped with a 337 nm nitrogen laser. An acceleration voltage of 25 kV was applied. The polymer sample was dissolved in THF with Milli-Q water containing 0.1% formic acid at a concentration of 0.8 mg mL−1. The matrix used was 2,5-dihydroxybenzoic acid (DHBA) (Pierce) and was dissolved in THF at a concentration of 30 mg mL−1.

3.4. General Procedure for the Depolymerization of Polylactide

The depolymerization reaction was carried out under an inert atmosphere. In a Braun LabMaster glovebox, a magnetically stirred reactor vessel (10 mL) was charged with polylactide. In a vial (4 mL), the metal complex was dissolved in MeOH or EtOH and added to the polymer. The reaction mixture was stirred at room temperature. At desired times, small aliquots of the reaction mixture were sampled, dissolved in CDCl3 or C6D6, and analyzed using 1H NMR spectroscopy. At the end of the depolymerization, the reaction was stopped with CH2Cl2 and dried under vacuum. The conversion of PLA, methyl lactate, and oligomers were calculated from 1H NMR, by the following equations:
Xint (%) = 1 − ([Int])/([Int]0) × 100
S Me-La (%) = ([Me-La])/([Int] − [Int]0) × 100
Y Me-La (%) = Xint × S_( Me-La)

3.5. General Procedure of PET Glycolysis

For each experiment, 0.200 g PET particles, the opportune volume of ethylene glycol, and a predicted amount of the catalyst were added to a 25 mL reaction tube with a magnetic stirrer. Glycolysis reactions were carried out at 180 °C for one hour. After this time, the reaction mixture was cooled to room temperature, and about 12 mL of distilled water was added; the resulting mixture was stirred and then filtered to eliminate the residual PET that was dried at 60 °C in vacuo to constant weight.
The water solution was concentrated and stored at 4 °C for the night. The crystals of BHET were recovered via filtration, dried, and weighted.
The conversion of PET was calculated by the following equation:
conv % = (W0 − Wr)/W0 × 100
where W0 is the initial weight of PET, and Wr is the weight of residual PET.
The water solution was concentrated by using a vacuum rotary evaporator at 70 °C and then refrigerated at 0 °C for 12 h to obtain white crystals of pure BHET.
The selectivity and yield of BHET were calculated according to the following equations:
Sel BHET % = (mol BHET crystals)/(mol PET soluble) × 100
Yield BHET % = (mol BHET crystals)/(PET initial) × 100

4. Conclusions

In this work, a study of the catalytic behavior of commercial metal amides in the chemical degradation of polylactide (PLA) and polyethylene terephthalate (PET) via alcoholysis was described.
In the degradation of PLA performed with methanol or ethanol at room temperature, the magnesium amide showed the highest activity, followed by Zn and Y amides, while the amides of group 4 metals were scarcely efficient. In the ethanolysis of PLA and the glycolysis of PET conducted at higher temperatures, the good activity of the Zn amide was preserved, while for Mg and Y amides, a significant decrease was observed. Surprisingly, for the glycolysis of PET, the Zr amide was revealed to be the best catalyst, enabling 78% BHET yield within one hour at 180 °C, which is a result that is comparable to the best ones described in the literature for metal catalysts. These studies represent the first examples of PET glycolysis mediated by homogeneous zirconium and yttrium catalysts.
The obtained results highlighted that simple non-toxic metal compounds can represent a valid tool to efficiently promote the sustainable recycling of exhausted polymeric materials into chemical products of synthetic interest.

Author Contributions

Investigation, F.S., R.C.L. and A.A.; conceptualization, M.L.; writing—review and editing, M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Patrizia Oliva for NMR assistance and Patrizia Iannece for MS spectrometry.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Millican, J.M.; Agarwal, S. Plastic Pollution: A Material Problem? Macromolecules 2021, 54, 4455–4469. [Google Scholar] [CrossRef]
  2. MacLeod, M.A.; Arp, H.P.H.; Tekman, M.B.; Jahnke, A. The global threat from plastic pollution. Science 2021, 373, 61–65. [Google Scholar] [CrossRef]
  3. Prata, J.C.; Silva, A.L.P.; da Costa, J.P.; Mouneyrac, C.; Walker, T.R.; Duarte, A.C.; Rocha-Santos, T. Solutions and Integrated Strategies for the Control and Mitigation of Plastic and Microplastic Pollution. Int. J. Environ. Res. Public Health 2019, 16, 2411. [Google Scholar] [CrossRef] [Green Version]
  4. Geyer, R.; Jambeck, J.R.; Law, K.L. Production, use, and fate of all plastics ever made. Sci. Adv. 2017, 3, e1700782. [Google Scholar] [CrossRef] [Green Version]
  5. Simon, N.; Raubenheimer, K.; Urho, N.; Unger, S.; Azoulay, D.; Farrelly, T.; Sousa, J.; van Asselt, H.; Carlini, G.; Sekomo, C.; et al. A binding global agreement to address the life cycle of plastics. Science 2021, 373, 43–47. [Google Scholar] [CrossRef]
  6. Schneiderman, D.K.; Hillmyer, M.A. 50th Anniversary Perspective: There Is a Great Future in Sustainable Polymers. Macromolecules 2017, 50, 3733–3749. [Google Scholar] [CrossRef]
  7. Zhu, Y.; Romain, C.; Williams, C.K. Sustainable Polymers from Renewable Resources. Nature 2016, 540, 354. [Google Scholar] [CrossRef]
  8. Filiciotto, L.; Rothenberg, G. Biodegradable Plastics: Standards, Policies, and Impacts. ChemSusChem 2021, 14, 56–72. [Google Scholar] [CrossRef]
  9. Kosloski-Oh, S.C.; Wood, Z.A.; Manjarrez, Y.; Pablo de los Rios, J.; Fieser, M.E. Catalytic methods for chemical recycling or upcycling of commercial polymers. Mater. Horiz. 2021, 8, 1084. [Google Scholar] [CrossRef] [PubMed]
  10. De Hoe, G.X.; Şucu, T.; Shaver, M.P. Sustainability and Polyesters: Beyond Metals and Monomers to Function and Fate. Acc. Chem. Res. 2022, 55, 1514–1523. [Google Scholar] [CrossRef]
  11. Zhang, X.; Fevre, M.; Jones, G.O.; Waymouth, R.M. Catalysis as an Enabling Science for Sustainable Polymers. Chem. Rev. 2018, 118, 839. [Google Scholar] [CrossRef]
  12. Schwach, G.; Coudane, J.; Engel, R.; Vert, M. Zn lactate as initiator of dl-lactide ring opening polymerization and comparison with Sn octoate. Polym. Bull. 1996, 37, 771–776. [Google Scholar] [CrossRef]
  13. Nijenhuis, A.J.; Grijpma, D.W.; Pennings, A.J. Lewis acid catalyzed polymerization of L-lactide. Kinetics and mechanism of the bulk polymerization. Macromolecules 1992, 25, 6419–6424. [Google Scholar] [CrossRef]
  14. Wang, X.; Liao, K.; Quan, D.; Wu, Q. Bulk Ring-Opening Polymerization of Lactides Initiated by Ferric Alkoxides. Macromolecules 2005, 38, 4611–4617. [Google Scholar] [CrossRef]
  15. Kricheldorf, H.R.; Berl, M.; Scharnagl, N. Poly(lactones). 9. Polymerization mechanism of metal alkoxide initiated polymerizations of lactide and various lactones. Macromolecules 1988, 21, 286–293. [Google Scholar] [CrossRef]
  16. Stevels, W.M.; Ankoné, M.J.K.; Dijkstra, P.J.; Feijen, J. Kinetics and Mechanism of l-Lactide Polymerization Using Two Different Yttrium Alkoxides as Initiators. Macromolecules 1996, 29, 6132–6138. [Google Scholar] [CrossRef]
  17. Ruelas-Leyva, P.J.; Picos-Corrales, A.L. Lactides and Lactones Yielding Eco-Friendly and Biocompatible Polymers by Metal-Catalyzed Ring-Opening Polymerization. Mini-Rev. Org. Chem. 2021, 18, 680–689. [Google Scholar] [CrossRef]
  18. Pretula, J.; Slomkowski, S.; Penczek, S. Polylactides-Methods of synthesis and characterization. Adv. Drug Deliv. Rev. 2016, 107, 3–16. [Google Scholar] [CrossRef]
  19. Pilone, A.; Press, K.; Goldberg, I.; Kol, M.; Mazzeo, M.; Lamberti, M. Gradient Isotactic Multiblock Polylactides from Aluminum Complexes of Chiral Salalen Ligands. J. Am. Chem. Soc. 2014, 136, 2940–2943. [Google Scholar] [CrossRef]
  20. Ovitt, T.M.; Coates, G.W. Stereochemistry of Lactide Polymerization with Chiral Catalysts: New Opportunities for Stereocontrol Using Polymer Exchange Mechanisms. J. Am. Chem. Soc. 2002, 124, 1316. [Google Scholar] [CrossRef]
  21. Osten, K.M.; Aluthge, D.C.; Mehrkhodavandi, P. Ligand Design in Enantioselective Ring-Opening Polymerization of Lactide. In Ligand Design in Metal Chemistry: Reactivity and Catalysis; Wiley: Hoboken, NJ, USA, 2016; pp. 270–307. [Google Scholar]
  22. Tschan, M.J.L.; Gauvin, R.M.; Thomas, C.M. Controlling polymer stereochemistry in ring-opening polymerization: A decade of advances shaping the future of biodegradable polyesters. Chem. Soc. Rev. 2021, 50, 13587–13608. [Google Scholar] [CrossRef] [PubMed]
  23. Bucknall, D.G. Plastics as a Materials System in a Circular Economy: Plastics in the Circular Economy. Philos. Trans. R. Soc. A 2020, 378, 20190268. [Google Scholar] [CrossRef] [PubMed]
  24. Coates, G.W.; Getzler, Y.D.Y.L. Chemical Recycling to Monomer for an Ideal, Circular Polymer Economy. Nat. Rev. Mater. 2020, 5, 501. [Google Scholar] [CrossRef]
  25. Payne, J.; McKeown, P.; Jones, M.D. A circular economy approach to plastic waste. Polym. Degrad. Stab. 2019, 165, 170–181. [Google Scholar] [CrossRef]
  26. Payne, J.; Jones, M.D. The Chemical Recycling of Polyesters for a Circular Plastics Economy: Challenges and Emerging Opportunities. ChemSusChem 2021, 14, 4041–4070. [Google Scholar] [CrossRef]
  27. Vogt, B.D.; Stokes, K.K.; Kumar, S.K. Why Is Recycling of Postconsumer Plastics so Challenging? ACS Appl. Polym. Mater. 2021, 3, 4325. [Google Scholar] [CrossRef]
  28. Available online: https://www.morganstanley.com.au/ideas/the-circular-economy-of-plastics (accessed on 5 September 2022).
  29. Santulli, F.; Lamberti, M.; Mazzeo, M. A Single Catalyst for Promoting Reverse Processes: Synthesis and Chemical Degradation of Polylactide. ChemSusChem 2021, 14, 5470–5475. [Google Scholar] [CrossRef]
  30. Santulli, F.; Gravina, G.; Lamberti, M.; Tedesco, C.; Mazzeo, M. Zinc and magnesium catalysts for the synthesis for PLA and its degradation: Clues for catalyst design. Mol. Catal. 2022, 528, 112480. [Google Scholar] [CrossRef]
  31. Payne, J.; McKeown, P.; Driscoll, O.; Kociok-Kohn, G.; Emanuelsson, E.A.C.; Jones, M.D. Make or break: Mg(II)- and Zn(II)-catalen complexes for PLA production and recycling of commodity polyesters. Polym. Chem. 2021, 12, 1086–1096. [Google Scholar] [CrossRef]
  32. Payne, J.M.; Kociok-Köhn, G.; Emanuelsson, E.A.C.; Jones, M.D. Zn(II)- and Mg(II)-Complexes of a Tridentate {ONN} Ligand: Application to Poly(lactic acid) Production and Chemical Upcycling of Polyesters. Macromolecules 2021, 54, 8453–8469. [Google Scholar] [CrossRef]
  33. Payne, J.M.; Kamran, M.; Davidson, M.G.; Jones, M.D. Versatile Chemical Recycling Strategies: Value-Added Chemicals from Polyester and Polycarbonate Waste. ChemSusChem 2022, 15, e202200255. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, R.; Xu, G.; Lv, C.; Dong, B.; Zhou, L.; Wang, Q. Zn(HMDS)2 as a Versatile Transesterification Catalyst for Polyesters Synthesis and Degradation toward a Circular Materials Economy Approach. ACS Sustain. Chem. Eng. 2020, 8, 18347–18353. [Google Scholar] [CrossRef]
  35. Yang, R.; Xu, G.; Dong, B.; Guo, X.; Wang, Q. Selective, Sequential, and “One-Pot” Depolymerization Strategies for Chemical Recycling of Commercial Plastics and Mixed Plastics. ACS Sustain. Chem. Eng. 2022, 10, 9860–9871. [Google Scholar] [CrossRef]
  36. Dutt, K.; Soni, R.K. A review on synthesis of value added products from polyethylene terephthalate (PET) waste. Polym. Sci. Ser. B 2013, 55, 430–452. [Google Scholar] [CrossRef]
  37. Fliedel, C.; Vila-Vicosa, D.; Calhorda, M.J.; Dagorne, S.; Aviles, T. Dinuclear Zinc-N-Heterocyclic Carbene Complexes for Either the Controlled Ring-Opening Polymerization of Lactide or the Controlled Degradation of Polylactide Under Mild Conditions. ChemCatChem 2014, 6, 1357–1367. [Google Scholar] [CrossRef]
  38. Pereira, C.S.M.; Silva, V.M.T.M.; Rodrigues, A.E. Ethyl lactate as a solvent: Properties, applications and production processes—A review. Green Chem. 2011, 13, 2658–2671. [Google Scholar] [CrossRef]
  39. Román-Ramírez, L.A.; Powders, M.; McKeown, P.; Jones, M.D.; Wood, J. Ethyl Lactate Production from the Catalytic Depolymerisation of Post-consumer Poly(lactic acid). J. Polym. Environ. 2020, 28, 2956–2964. [Google Scholar] [CrossRef]
  40. Barnard, E.; Arias, J.J.R.; Thielemans, W. Chemolytic depolymerisation of PET: A review. Green Chem. 2021, 23, 3765. [Google Scholar] [CrossRef]
  41. Damayanti, N.; Wu, H.S. Strategic possibility routes of recycled PET. Polymers 2021, 13, 1475. [Google Scholar] [CrossRef]
  42. Fukushima, K.; Coady, D.J.; Jones, G.O.; Almegren, H.A.; Alabdulrahman, A.M.; Alsewailem, F.D.; Horn, H.W.; Rice, J.E.; Hedrick, J.L. Unexpected efficiency of cyclic amidine catalysts in depolymerizing poly(ethylene terephthalate). J. Polym. Sci. Part A Polym. Chem. 2013, 51, 1606–1611. [Google Scholar] [CrossRef]
  43. Jehanno, C.; Demarteau, J.; Mantione, D.; Arno, M.C.; Ruipérez, F.; Hedrick, J.L.; Dove, A.P.; Sardon, H. Selective Chemical Upcycling of Mixed Plastics Guided by a Thermally Stable Organocatalyst. Angew. Chem. Int. Ed. 2021, 133, 6784. [Google Scholar] [CrossRef]
  44. Jehanno, C.; Perez-Madrigal, M.M.; Demarteau, J.; Sardon, H.; Dove, A.P. Organocatalysis for depolymerisation. Polym. Chem. 2019, 10, 172–186. [Google Scholar] [CrossRef] [Green Version]
  45. Delle Chiaie, K.R.; McMahon, F.R.; Williams, E.J.; Price, M.J.; Dove, A.P. Dual-catalytic depolymerization of polyethylene terephthalate (PET). Polym. Chem. 2020, 11, 1450–1453. [Google Scholar] [CrossRef]
Scheme 1. Metal amides screened in this work.
Scheme 1. Metal amides screened in this work.
Catalysts 12 01193 sch001
Scheme 2. Depolymerization of PLA via alcoholysis.
Scheme 2. Depolymerization of PLA via alcoholysis.
Catalysts 12 01193 sch002
Figure 1. 1H NMR (CDCl3, 400 MHz) monitoring of PLA methanolysis with assignment of internal, chain-end, and methyl lactate methine groups (run 3, Table 1).
Figure 1. 1H NMR (CDCl3, 400 MHz) monitoring of PLA methanolysis with assignment of internal, chain-end, and methyl lactate methine groups (run 3, Table 1).
Catalysts 12 01193 g001
Figure 2. 1H NMR (CDCl3, 400 MHz) monitoring of PLA ethanolysis with assignment of internal (black), chain-end groups (blue and green), and ethyl lactate (red) alkyl protons.
Figure 2. 1H NMR (CDCl3, 400 MHz) monitoring of PLA ethanolysis with assignment of internal (black), chain-end groups (blue and green), and ethyl lactate (red) alkyl protons.
Catalysts 12 01193 g002
Figure 3. Yield of ethyl lactate by Zr and Y amides at different temperatures.
Figure 3. Yield of ethyl lactate by Zr and Y amides at different temperatures.
Catalysts 12 01193 g003
Scheme 3. Glycolysis of PET with production of BHET.
Scheme 3. Glycolysis of PET with production of BHET.
Catalysts 12 01193 sch003
Figure 4. 1H NMR (CDCl3, 400 MHz, 25 °C) BHET obtained via crystallization from cold water.
Figure 4. 1H NMR (CDCl3, 400 MHz, 25 °C) BHET obtained via crystallization from cold water.
Catalysts 12 01193 g004
Figure 5. ESI-MS spectra of BHET obtained via crystallization from cold water.
Figure 5. ESI-MS spectra of BHET obtained via crystallization from cold water.
Catalysts 12 01193 g005
Table 1. Methanolysis of PLLA using metal amides.
Table 1. Methanolysis of PLLA using metal amides.
Run [a]CatTime
(h)
XInt [b]
(%)
SMe-La [b]
(%)
YMe-La [b]
(%)
1Zn23910039
2Mg29010090
3Y199810098
4Ti241008989
5Zr811698
6Hf210525
[a] All reactions were carried out by using 72 mg of PLA and 10 µmol of catalyst (1 mol% relative to ester linkages) at 25 °C with 1 mL of MeOH. [b] Determined by 1H NMR.
Table 2. Ethanolysis of PLLA using metal amides.
Table 2. Ethanolysis of PLLA using metal amides.
Run [a]CatT
(°C)
XInt [b]
(%)
SEt-La [b]
(%)
YEt-La [b]
(%)
1Zn251008383
2Mg25939891
3Y251009191
4Ti25753929
5Zr253512
6Hf2550.21
7Zn80100100100
8Mg80575029
9Y801710017
10Ti80100100100
11Zr80100100100
12Hf80895448
13Y60575833
14Zr60416828
[a] All reactions were carried out by using 500 mg of PLLA sample and 7 × 10−5 mol of catalyst (1 mol% relative to ester linkages) with 10 mL of EtOH for 24 h. [b] Determined by 1H NMR.
Table 3. Glycolysis of PET under solvent-free conditions at 180 °C.
Table 3. Glycolysis of PET under solvent-free conditions at 180 °C.
Run [a]CatConv
(%)
Sel BHET
(%)
Yield BHET
(%)
1Zn922624
2Mg962422
3Y965826
4Ti52158
5Zr867865
6Hf732821
[a] Reaction conditions: 200 mg of PET bottle (Mn ≈ 42.000 g mol−1), 0.8 mL of EG (27.8 equivalents relative to the ester bonds), and catalyst (0.013 equivalents relative to the ester bonds) at 180 °C for 1 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Santulli, F.; Lamberti, M.; Annunziata, A.; Lastra, R.C.; Mazzeo, M. The Contribution of Commercial Metal Amides to the Chemical Recycling of Waste Polyesters. Catalysts 2022, 12, 1193. https://doi.org/10.3390/catal12101193

AMA Style

Santulli F, Lamberti M, Annunziata A, Lastra RC, Mazzeo M. The Contribution of Commercial Metal Amides to the Chemical Recycling of Waste Polyesters. Catalysts. 2022; 12(10):1193. https://doi.org/10.3390/catal12101193

Chicago/Turabian Style

Santulli, Federica, Marina Lamberti, Andrea Annunziata, Rita Chiara Lastra, and Mina Mazzeo. 2022. "The Contribution of Commercial Metal Amides to the Chemical Recycling of Waste Polyesters" Catalysts 12, no. 10: 1193. https://doi.org/10.3390/catal12101193

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop