Open Access Article
Deepthi Thomas
ab,
Parvathy C.a and
Benny K. George
*a
aVikram Sarabhai Space Centre, Indian Space Research Organization, Thiruvananthapuram 695022, Kerala, India. E-mail: bkgeorge63@gmail.com
bDepartment of Applied Chemistry, Cochin University of Science and Technology, Cochin 682022, India
First published on 5th November 2025
In spite of its many applications and desirable properties, the non-biodegradable nature of polyethylene terephthalate (PET) raises concerns about the accumulation of post-consumer products. Recycling these end-of-life polymers is the most effective solution to address this issue. Mechanical recycling, the primary commercial recycling method for PET, often produces materials with inferior properties, and hence chemical recycling, or tertiary recycling, is considered a viable alternative. Glycolysis, which is a transesterification reaction in which the PET molecule is depolymerized to its monomer, bis(hydroxyethyl) terephthalate (BHET), has gained significant commercial interest due to its mild reaction conditions. In the present work, magnetic cobalt oxide nanoparticles (CONP) were synthesized using different methods: NaBH4 reduction route, hydrothermal route, and combustion route, and experimented as catalysts for the glycolysis of PET. The catalysts produced were characterized using FTIR, SEM-EDS, surface area and XRD techniques. Glycolysis of PET was performed with the CONP catalysts obtained from each synthesis method and the resulting BHET was analysed using DSC, HPLC, and NMR techniques. Among the catalysts, CONP prepared via the NaBH4 route showed the best performance, achieving a BHET yield of 97% with just 1% catalyst at 180 °C within a reaction time of 2 hours. The effects of various reaction conditions, including temperature, reaction time, and the PET/EG ratio, were also investigated. Notably, the CONP–NaBH4 catalyst is magnetically separable from the reaction mixture after the process and can be reused multiple times without considerable loss in efficiency. CONP–NaBH4 is structurally distinct from other cobalt oxides, featuring a core–shell structure with a nanocrystalline cobalt boride core and an amorphous cobalt oxide, oxyhydroxide, hydroxide shell. This was established by STEM, HRTEM, XPS, Raman, and FTIR analyses. The amorphous cobalt oxide shell drives catalytic activity, while the cobalt boride core imparts ferromagnetism. Superior catalytic performance is attributed to its high surface area, porosity, and presence of surface hydroxyl groups.
Polyethylene terephthalate (PET), a polyester made from terephthalic acid and ethylene glycol is utilized in various fields, including food packaging and aerospace. However, PET's non-biodegradable nature raises concerns about the accumulation of post-consumer PET products in the environment. Recycling end-of-life polymers is a viable solution to address this issue. Recycling strategies reported for PET waste are re-extrusion (primary recycling), mechanical recycling (2° recycling), and tertiary or feedstock recycling. Though mechanical recycling is the major route of recycling, it always results in a downgrade product.1,2 Among the various methods for tertiary recycling of PET, glycolysis has gained commercial interest due to its favourable reaction conditions. Glycolysis is a transesterification reaction that depolymerizes the PET molecule using ethylene glycol to produce bis(hydroxyethyl) terephthalate (BHET). BHET can be used as a starting material for PET polymerisation or to produce other value-added products like polyurethane or acrylates.1
PET glycolysis is a slow process without a catalyst. The catalysts used for PET glycolysis fall into several categories as: metal derivatives,3–6 ionic liquids (IL),7–10 deep eutectic solvents (DES),11–13 and organic catalysts.14,15 Metal-based catalysts include metal salts,16,17 metal oxides,6,18 metal–organic frameworks (MOFs),19 metal nanoparticles,20 metal oxide-doped graphene,21 and carbon nanotubes (CNTs),22 and layered double hydroxides23,24 (LDH). Nano catalysts have gained attention due to their unique properties, which enhance the catalytic process.
Cobalt-based salts, mixed metal oxides, and cobalt nanoparticles have been identified as effective catalysts for achieving high yields of BHET. For instance, Chen et al. reported that cobalt–aluminium mixed oxides produced a BHET yield of 69%.25 Ultra small cobalt nanoparticles were found to function as regenerable catalysts for the precipitation of BHET from PET without the need for external solvents.20 Another promising catalyst is CoAlCO3-LDH, which demonstrated a remarkable BHET yield of 96%. Cobalt chloride and cobalt-based metal–organic complexes,5 CoFe2O4 and its composites, exhibited good catalytic activity.26 Notably, a cobalt-based ionic liquid grafted onto graphene exhibited high catalytic activity, achieving a 95% yield of BHET under mild conditions, along with simple recovery methods.27
Cobalt oxide and mixed oxide spinels are reported as glycolysis catalysts by Imran et al.3 Although cobalt oxide is known to act as a catalyst in glycolysis, none of its reported forms possess inherent magnetic properties that would allow for magnetic separation. The cobalt oxide spinels reported by Imran et al. were prepared by precipitation of cobalt hydroxide from corresponding nitrates, followed by calcination at high temperature to get the oxide.
In the present work, magnetic cobalt nanoparticles (CONP) were prepared by three different routes, viz.; NaBH4 routes,28 hydrothermal route,29 and combustion routes30 and demonstrated as catalysts for the glycolysis of PET. The prepared catalysts were characterized using FTIR, SEM-EDS, surface area analysis, and XRD techniques. PET glycolysis was performed using CONP catalysts prepared by different synthetic routes. The BHET produced was characterized using DSC, HPLC, and NMR techniques. Among the catalysts tested, the CONP catalyst prepared via the NaBH4 route proved to be the most effective, yielding 97% BHET with just 1% catalyst at a temperature of 180 °C within a reaction time of 2 hours. Additionally, the effects of various reaction conditions, such as temperature, reaction time, and the ratio of PET to EG, were also investigated. The CONP–NaBH4 catalyst is magnetically separable from the solution after the reaction and can be reused without losing its effectiveness.
Synthesis by modified solution combustion method:30 aqueous solutions of cobalt nitrate hexahydrate (3.305 g) and sucrose (3.805 g) were dissolved in 40 mL of distilled water. The solution mixture was magnetically stirred for 1 h and heated to 350 °C in air. After complete water evaporation, auto ignition started with the release of large amounts of gases. Finally, a fine black product was collected, washed with water, dried, and calcined at 600 °C for 5 h. The final product is labelled as CONP-combustion.
Hydrothermal method:29 in a typical synthesis, 1.475 g of cobalt chloride (CoCl2) and 0.305 g of urea (CO(NH2)2) were dissolved in 25 mL of distilled water to form homogeneous solutions, respectively. The urea solution was then added dropwise to the CoCl2 solution with stirring. The mixture was then transferred to a 100 mL Teflon-liner autoclave, which was sealed and heated to 105 °C for 6 h. After the autoclave was cooled to room temperature, the resulting pink precipitate was separated by centrifugation, washed three times with distilled water and ethanol, respectively, and dried in an oven at 70 °C. The dried product was then sintered at 300 °C for 3 hours. The black powder obtained is named as CONP-hydrothermal.
Schematic of the above-mentioned syntheses are shown in Fig. 1.
PET Glycolysis:16,24 1.0 g of PET pieces obtained from mineral water bottles, 0.01 g (1%) CONP, and 10 ml ethylene glycol were heated in a 100 ml two-necked round-bottom flask fitted with a water jacketed condenser at 180 °C for 2 hours with a stirring rate of 400 rpm at ambient pressure. A heating mantle with temperature control was used for the experiment. Temperature of the reaction mixture is also monitored using a thermometer inserted into the reaction mixture. After 2 hours, the temperature was decreased to 100 °C, and un-depolymerized PET that remained in the solution was separated. Then, hot water was added to the system and filtered to separate the catalyst. The filtrate was kept at 4 °C for 12 hours to precipitate BHET. Precipitated BHET was washed with water and dried at 80 °C for 12 hrs. Dried BHET was characterized by FTIR, FTNMR, DSC, and HPLC techniques.
PET conversion was calculated using the following equation.
The yield of BHET was determined through HPLC analysis. First, the products from the glycolysis reaction were filtered to remove the catalyst and any insoluble oligomers. The resulting filtrate was then diluted to a final volume of 100 ml using HPLC-grade methanol. 100 μL of this solution was further diluted to 10 ml with methanol before injection into the HPLC system. Using a standard BHET, a calibration graph was plotted based on the peak areas obtained from the HPLC analysis. The concentration of BHET in the sample was then estimated using this calibration graph. The yield of BHET was calculated as follows:
:
20). Scanning Transmission Electron Microscopic (STEM) analysis was carried out using GeminiSEM 500 (Make: Carl Zeiss). Sample was dispersed in IPA and drop casted on copper grid. Analysis of the dried sample was done using electron beam of accelerating voltage 20V and imaging was done by STEM detector. High resolution transmission electron microscopy (HRTEM) analysis was done using Geol JEM 2100 transmission electron microscope with a 200 kV electron gun.
The BET surface area of the samples was determined using Quantachrom, Novtouch, lLX2. Degassing of the sample was done at 150 °C for 6 hours in flowing nitrogen gas before the sample analysis. Pore size distribution is computed by DFT method. Pore volume is calculated at partial pressure 0.99. The melting point was determined using a TA Instruments 2920 differential scanning calorimeter (DSC). The sample was placed in an aluminum pan and heated from 25 °C to 250 °C at a rate of 5 °C per minute under a nitrogen flow of 10 mL per minute. Both 13C and 1H nuclear magnetic resonance (NMR) spectra were recorded using a Bruker Avance 400 MHz NMR spectrometer, with CDCl3 used as the solvent. X-ray diffraction (XRD) measurements were conducted using a Bruker D8 Discover X-ray diffractometer operating with a copper anode (40 kV, 40 mA). Elemental analysis was performed by inductively coupled plasma atomic emission spectrometry (ICP-AES) using a Thermofisher Scientific iCAP 7400. The magnetic characterization was performed using a Quantum design VersaLab in a field of −30 kOe to 30 kOe at 300K. XPS analysis was caried out with Thermo Fisher Scentific Nexsa G2 surface analysis system. High-performance liquid chromatography (HPLC) analysis was performed using a PerkinElmer UHPLC system (Model: Flexar F-10). Chromatographic separation was achieved employing a Brownlee analytical C8 column (100 mm × 4.6 mm), with methanol serving as the mobile phase. The flow rate was maintained at 1 mL min−1 and the injection volume was set to 5 μL. Eluted analytes were detected using a UV detector at a wavelength of 254 nm.
Average BET surface areas of the prepared catalysts with standard deviations are listed in Table 1. Surface area is highest for catalyst prepared through sodium borohydride route and lowest for the one prepared through hydrothermal route. Total pore volume calculated by DFT is shown in Table 1. Pore distribution curves are given in Fig. S4. All the three samples are mesoporous with pores size between 2 and 20 nm. Fig. 2a–c show the adsorption–desorption hysteresis loops of CONPs, which are characteristic of mesoporous materials.
| Catalyst | Surface area (m2 g−1) | Total pore volume (cm3 g−1) | |
|---|---|---|---|
| 1 | CONP–NaBH4 | 79.1 ± 12.1 | 0.253 |
| 2 | CONP-hydrothermal | 1.9 ± 0.99 | 0.008 |
| 3 | CONP-combustion | 29.5 ± 5.8 | 0.052 |
Overlaid FTIR spectra of the CONP are shown in Fig. 3. Cobalt oxide prepared via hydrothermal and combustion routes showed sharp peaks at 651 cm−1 and 600 cm−1 due to Co3O4 spinels. FTIR spectral peak of CONP prepared by borohydride reduction indicates the amorphous nature. In addition to peaks around 650 and 592 cm−1, peaks were observed at 1420 and 1048 cm−1 due to –OH deformation modes of –CoO(OH).31 Borate stretching falls in the range –BO3 at 1433 cm−1 and –BO4 structures at 1102 cm−1.32 Stretching and bending vibrations of adsorbed moisture are seen around 3400 and 1630 cm−1.
Overlaid Raman spectra of the CONP are shown in Fig. 4. Peaks at, 482, 517 and 687 cm−1 correspond to Eg, F2g and A1g modes of the Co3O4 respectively.33–35 Raman peaks of CONP–NaBH4 are broad compared to the other two indicating amorphous nature of the catalyst. Difference in crystallinity of the catalysts are evident from the FWHM of the peak at 687 cm−1. FWHM of the peak is 9.1 cm−1 for CONP–NaBH4, while it is 5.7 and 6.3 cm−1 for CONP-combustion and CONP-hydrothermal respectively.
XRD patterns of the three catalysts are shown in Fig. 5. The diffraction pattern is characteristic of Cubic Co3O4 for CONP-hydrothermal and CONP-combustion, while no diffraction pattern obtained for CONP–NaBH4 confirming its complete amorphous nature. The surface morphology of the prepared catalysts was studied using FESEM images (Fig. 6). CONP-hydrothermal showed well defined rod-shaped nano aggregates of micrometre length. Size of each particle in aggregate is ≈20 nm. CONP-combustion showed aggregates with particles of nanometre to sub-micrometre size, with micropores in between. SEM image of CONP–NaBH4 shows a foam like morphology. SEM image indicates that material is nanoporous.
![]() | ||
| Fig. 6 FESEM images of the CONP catalysts prepared (a) CONP hydrothermal, (b) CONP combustion and (c) CONP–NaBH4. | ||
There are different views among the scientific community about the composition of products formed during hydrolysis/ethanolysis by sodium borohydride in the presence of cobalt chloride. While the evolution of hydrogen is commonly agreed upon, what happens to the cobalt chloride during hydrolysis/alcoholysis is a question of debate. Many studies consider the black precipitate as cobalt borides (CoxB) and some reports say a mixture of cobalt borides and cobalt metal.36,37 Based on our characterization, especially Raman spectroscopy, the black precipitate formed is mainly cobalt oxide. ICP-AES analysis of CONP–NaBH4 (Table S1) showed the presence of 4.7% B, 44% cobalt, and 1.7% sodium in the sample. The estimated boron content corresponds to 30% of CoB (calculation S1). Sodium present in the sample suggests that this percentage could be even lower, as some fraction of boron may be present as borate. Borate vibrations are present in the FTIR spectrum of CONP–NaBH4, but the peak at 789 cm−1 corresponding to the ν(Co–B) in cobalt boride is absent in the FTIR spectrum.38
XPS analysis of CONP–NaBH4 was carried out get further clarity on the structure (Fig. S6). The XPS spectra of Co 2p level shows characteristic binding energies of Co2+ and Co3+ oxidation states correspond to the presence of Co3O4, CoO and Co (OH)2.36,37 B 1s XPS spectrum showed as single peak at 191 eV due the presence of BO2− species,36,39 O 1s XPS revealed the presence of several non-equivalent states of oxygen atoms. The binding energy values correspond to O2−, –OH, O2− spinel, and H2O.40,41
The STEM image of the catalyst revealed a core shell morphology as shown in Fig. 7. The shell thickness is measured as 14–16 nm from the STEM image. Netskina et al.42 have proposed a core shell structure for amorphous ferromagnetic cobalt–boron composition reduced by sodium borohydride. The shell thickness reported by them is 2–3 nm, but the catalyst reported here shows a shell thickness of 14–16 nm. The higher shell thickness may be the reason for absence of CoB peak in XPS. Similar observation was reported in the XPS of amorphous cobalt borate nanosheet-coated cobalt boride.43
The HRTEM images of CONP–NaBH4 is shown in Fig. 8. Core–shell structure is evident in TEM images also. The core showing nanocrystalline nature with a lattice fringe spacing of 0.23 nm corresponding to d spacing of 101 plane of CoB (JCPDS card 03-0959) and shell is amorphous in nature.44 The SAED pattern shows diffuse rings rather than sharp spots, indicating the catalyst is nanocrystalline.
:
EG ratio of 1
:
5. BHET crystallized out from glycolysate was characterized by NMR and FTIR spectroscopy (Fig. 9A, C and D). FTIR spectrum of the separated glycolysis product has characteristic peaks of BHET. DSC of the BHET is shown in Fig. 9B. The purity of the separated BHET is evident from the sharp endotherm, with peak onset at 109 °C due to the melting of BHET. The reported melting point of BHET dimer is between 162 to 166 °C. The obtained melting point ‘109 °C’ is matching with that reported for pure BHET. The peaks and integral values in 1H NMR spectrum clearly match with BHET structure. The signals at 61.3, 67.0, 166.0, 133.9, and 129.7 ppm in 13C NMR are also characteristic of the BHET monomer. Non-appearance of peak at 63 ppm in 13C NMR and 5 ppm in 1H NMR confirms the absence of dimer. It is established from NMR analysis that the separated BHET is 100% pure as the spectra do not show any additional peak.
![]() | ||
| Fig. 9 Characterisation of BHET (A) FTIR spectrum, (B) DSC, (C) 1H NMR spectrum and (D) 13C NMR spectrum. | ||
The glycolysate was analysed by HPLC to find out the BHET yield. Calibration graph for estimation of BHET yield is given in Fig. S7. Overlaid chromatograms of BHET standard and the PET glycolysates are shown in Fig. 10. Peak at 1.1 minutes is due to BHET, and no dimer peak (which is expected around 1.34 minutes) is present in any of the chromatograms.
![]() | ||
| Fig. 10 Overlaid HPLC of PET glycolysates obtained using different CONP catalysts and BHET standard solution. | ||
The PET conversion and BHET yield obtained with 1% of the CONP catalysts at an EG/PET ratio of 10, and a reaction temperature of 180 °C for 2 hours are given in the Fig. 11.
![]() | ||
| Fig. 11 BHET yield and PET conversion obtained using CONP catalysts (experimental conditions: reaction time 2H, catalyst 1%, EG/PET 10, temperature 180 °C). | ||
Among the catalysts prepared, the CONP synthesized using the sodium borohydride route demonstrated the best performance, achieving 97% yield of BHET and 100% conversion of PET. The second-best performer was the CONP-combustion method, which resulted in 83% yield of BHET and 86% conversion of PET. In contrast, performance of CONP produced via the hydrothermal route was the lowest, yielding only 70% BHET and 76% conversion of PET. Mass balance summary for glycolysis reactions with three catalysts is given in Table S3.
This is the first report on magnetically separable CONP prepared by NaBH4 reduction route for PET glycolysis. Amorphous ferromagnetic cobalt–boron composition reduced by sodium borohydride was studied widely as hydrogen generation catalysts,36,37,42 but their use as transesterification catalyst for PET has not explored yet. Though NaBH4 reduction was reported to make ultrasmall cobalt nanoparticles for PET glycolysis, their synthesis method is different20 and magnetic separation was not attempted. Performance of CONP–NaBH4 is in parr with the reported cobalt-based catalysts for PET glycolysis (Table 2). The highest BHET yield reported among cobalt based catalyst for PET glycolysis 99% is for CoAl LDH@Fe3O4.23 But CoAl LDH is not inherently magnetic and magnetic support is used to make it magnetically regenerable.
| Catalyst | BHET % | EG/PET | Time (h) | Catalyst (%) | Glycolysis temperature | Regeneration | |
|---|---|---|---|---|---|---|---|
| a BHET yield under optimum conditions are compared. | |||||||
| 1 | CoAl mixed oxides25 | 69 | 5 | 0.83 | 1 | 196 | Not reported |
| 2 | Ultrasmall cobalt nanoparticles20 | 77 | 20 | 3 | 1.5 | 180 | Regenerable by filtration |
| 3 | CoFe2O4 (ref. 46) | 77 | 6 | 6 | 4 | 190 | Magnetic regeneration |
| 4 | CoFe2O4@ZIF-8/ZIF-67 46 | 84 | 5 | 1 | 1 | 200 | |
| 5 | CoFe2O4/C10-OAC46 | 95 | 5 | 2.5 | 2 | 195 | |
| 6 | CoCl2 (anh)5 | 65 | 10 | 3 | 1 | 190 | Not reported |
| 7 | CoCl2 (anh)/dcype5 | 71 | |||||
| 8 | CoCl2 (anh)/dppe5 | 53 | |||||
| 9 | CoCl2 (anh)/dppf5 | 31 | |||||
| 10 | CoAlCO3-LDH@ Fe3O4 (ref. 23) | 99 | 10 | 2 | 1 | 180 | Magnetic regeneration |
| 11 | Cobalt-based ionic liquid grafted on graphene27 | 95 | 190 | ||||
| 12 | Cobalt oxides recycled from spent batteries47 | 10 | 8 | 2 | 1 | 196 | Not reported |
| 13 | Cobalt oxide–NaBH4 reduction (present work) | 97 | 5 | 2 | 1 | 180 | Magnetic regeneration |
Turnover Frequency (TOF) is a measure of catalytic activity and it represents the number of reactant molecules converted to product per active site on the catalyst per unit time. The TOF calculated for CONP–NaBH4 is 271 h−1, which is higher than reported for the commercial glycolysis catalyst Zn (OAc)2
45 (calculation S2).
Hydrogen gas is evolved as a byproduct in the preparation of CONP–NaBH4. Cobalt chloride is a well-established catalyst for hydrogen generation from NaBH4 through hydrolysis or alcoholysis reactions, and the catalyst CONP–NaBH4 could also be produced in the future as a valuable byproduct during hydrogen production. This approach presents an opportunity to combine clean energy generation with the principles of a circular economy for plastic manufacturing, aligning resource efficiency with sustainability goals.
Environmental footprint for the synthesis of CONP–NaBH4 is low compared to hydrothermal or combustion routes. Combustion routes and hydrothermal routes require temperatures up to 600 °C and 300 °C, respectively. Greenhouse gas (carbon dioxide) evolution is another drawback of hydrothermal and combustion routes.
EG/PET ratio was optimized by varying the ratio from 5 to 20 (Fig. 12B). Temperature, reaction time, and catalyst concentrations were fixed as 180 °C, 2 hours, and 1% respectively. BHET yield increased with the increase in EG/PET ratio and reached a maximum at a ratio of 10. Further increase in EG decreases the BHET yield. A similar trend is reported for cobalt aluminium LDH catalyst.23
Glycolysis increases with rising temperatures, reaching an optimal value before declining (Fig. 12C). The optimal reaction temperature for CONP–NaBH4 was identified as 180 °C; however, further increasing the temperature to 190 °C resulted in a decrease in BHET yield to 88%. It has been reported that high reaction temperatures may also promote the polymerization of the BHET formed.48
![]() | ||
| Fig. 13 (a) and (b) Magnetic separation of the catalyst from EG and from the reaction mixture after completion of glycolysis and (c) BHET yield during four regeneration cycles. | ||
Overlaid Raman spectra of the fresh catalyst and the catalyst separated after regeneration are shown in Fig. 14a. Peak postions are corresponding to F2g, Eg, and A1g phonon modes of Co3O4.34,35 Peak broadening and peak red shift are noticed compared to the spectrum before recycling. FWHM of the peak at 678 cm−1, was 9.1 cm−1 before regeneration, while it increased to 24.5 cm−1 for the regenerated catalyst. Increase in FWHM points towards the dissociation of nanoaggregates to smaller nanoparticles after glycolysis. Raman spectra of the catalyst were acquired after each cycle (Fig. S9) and the spectra are characteristic of cobalt oxide.
![]() | ||
| Fig. 14 Overlaid (a) Raman spectrum, (b) FTIR spectrum and (c) SEM image of CONP–NaBH4 as such and after regeneration. | ||
FTIR spectrum of the catalyst before glycolysis and after glycolysis is shown Fig. 14b. Catalyst after glycolysis also showed peaks corresponding to borate and cobalt oxide, like that of the initial catalyst. A small peak at 790 cm−1, which was not seen initially, is evident in the catalyst after glycolysis. As per the literature, this could be due to the Co–B stretching. This peak was not visible in the initial phase may be due to overlapping with borate peaks. SEM images of the catalyst confirm that the regenerated catalyst retains its morphology after reaction. EDS analysis of the regenerated catalyst showed a decrease in sodium compared to the original may be due to the dissolution of soluble sodium borate during reaction and washing. XPS analysis of the regenerated catalyst shows a spectrum similar to the original, but boron content is reduced after glycolysis, confirming the EDS inferences.
ICP-AES analysis of the filtrates from each step was performed to evaluate catalyst leaching into the solution. Cobalt concentration was consistently below 20 ppm (Table S3), confirming minimal leaching and demonstrating the effective separation of the catalyst. In addition, ICP-AES analysis of the recovered BHET failed to detect any cobalt content.
The hot filtration test was carried out to verify heterogeneous catalysis. Details of the test and calculation are given in SI (Tables S4 and S5). After 1 hour, the catalyst was filtered from the reaction mixture, and the reaction was allowed to proceed for another hour without the catalyst. The BHET yield after 2 hours dropped to 84%, compared to 98% in the presence of catalyst, confirming the heterogeneous catalytic pathway.
| This journal is © The Royal Society of Chemistry 2025 |