Songlei
Mo
,
Yong
Guo
,
Xiaohui
Liu
and
Yanqin
Wang
*
Shanghai Key Laboratory of Functional Materials Chemistry, Research Institute of Industrial Catalysis, School of Chemistry and Molecular Engineering, East China University of Science and Technology, Shanghai 200237, P. R. China. E-mail: wangyanqin@ecust.edu.cn
First published on 25th October 2023
The glycolysis of polyethylene terephthalate (PET) is considered as the most promising PET recycling strategy due to its mild reaction conditions and circularity. Herein, several metal-doped SBA-15 (M/SBA-15) as heterogeneous Lewis acid catalysts were investigated to screen the best catalyst and to understand the essence for their differences in PET glycolysis. It is found that the catalytic activity is positively correlated with the acid amount in different metal-doped SBA-15 and Ti/SBA-15 with different Ti loadings, with the exception of Nb/SBA-15. 4Ti/SBA-15 with abundant Lewis acid sites shows excellent PET glycolysis ability, with the yield of bis(2-hydroxyethyl) terephthalate (BHET) reaching 87.2% within 45 min at 190 °C (isolated yield 73%). While 4Nb/SBA-15 has a similar acid amount as that in 4Ti/SBA-15, but it has very low catalytic activity probably due to its strong oxygen affinity. Acetone–methanol-adsorption DRIFTS confirms its strong oxygen affinity, which would poison the acid sites by ethylene glycol (EG) and result in very low activity. This study demonstrates that PET glycolysis activity is really correlated with the amount of Lewis acid sites, but species with very strong oxygen affinity would lower its activity. Further diffuse reflection ultraviolet-visible (DR UV-vis) spectra analysis confirms that tetrahedral Ti4+ species in Ti/SBA-15 is the real active center. In addition, 4Ti/SBA-15 has good reusability and can be used in common real PET (dyed polyester fabric, PET bottles) glycolysis. These results indicate that increasing the amount of Lewis acid sites is a promising way to enhance PET glycolysis and 4Ti/SBA-15 with high activity and stability is a potential catalyst for industrial application.
Currently, numerous heterogeneous catalysts for PET glycolysis, including heterogeneous basic catalysts (e.g., Ca- and Mg-based catalysts)18–20 and Lewis acid catalysts (e.g., Zn-, Co-, Fe-, and Mn-based catalysts)12–15 have been developed. Compared with homogeneous catalysts, heterogeneous catalysts are easy to separate from the product and have less influence on the purity of BHET. Scheme 1 shows the mechanisms of depolymerizing PET catalyzed by basic catalysts and Lewis acid catalysts. The basic catalysts can activate EG through the deprotonation of the OH group by the Brønsted base sites to form catalytically active (and more nucleophilic) EG− species.6,13 Meanwhile, the Lewis acid catalysts can activate the CO bonds in the PET chains via the Lewis acid sites to make the electrons of CO bonds more biased toward oxygen.13,14 The carbon atoms are then susceptible to being attacked by EG to achieve the transesterification reaction.
Scheme 1 Reaction mechanisms of PET glycolysis with (a) basic catalysts and (b) Lewis acid catalysts. |
According to the literatures reported so far, the heterogeneous basic catalysts generally suffered from low reactivity and low yield of BHET (≤81%);18–20 thus, heterogeneous Lewis acid catalysts have been extensively studied. Wu et al.15 synthesized spinel-type MFe2O4 (MCo, Ni, Cu, and Zn) to depolymerize PET via glycolysis and found that the glycolysis activity was positively correlated with Lewis acid strength. The stronger Lewis acid sites interacted more strongly with the CO bonds in the PET chains, resulting in higher catalytic efficiency. Unfortunately, these catalysts were less active and took 6 h for total PET glycolysis, probably due to the less acid amounts of these metal composite oxides. Rinaldi and co-workers14 prepared ultra-small cobalt nanoparticles modified with tannic acid, which made the catalyst highly dispersed in EG. The PET conversion and BHET isolated yield reached 96% and 77% at 180 °C within 3 h, respectively. Similarly, Wang et al.12 synthesized Fe3O4 nano-dispersions modified with sodium citrate dihydrate for PET glycolysis. With this catalyst, PET can be completely depolymerized at 210 °C within 30 min, and the BHET yield reached 93%. All these works emphasized the importance of effective contact between the catalyst and PET, which seriously affected the depolymerization rate in PET glycolysis.
In addition to preparing ultra-small nanoparticles catalysts and modifying their surfaces to improve their dispersibility in EG, mesoporous catalysts with high specific surface area and rich pore structure would be another choice. The obvious advantage of this kind of catalyst is the easy separation compared to the nanoparticle catalyst. Therefore, in this work, the siliceous SBA-15, which has high specific surface area, rich pore structure, and is chemically inert, was selected as the catalyst support. In order to obtain an efficient PET glycolysis catalyst, we screened a series of representative metal ions (Nb,26 Ti,27 Zn,28,29 Sn (ref. 30–32)) with Lewis acidity and doped them in SBA-15 to depolymerize PET. It is important to note here that after doping with different metal ions, the textural properties and acid strength of these catalysts are similar. A positive correlation between catalytic activity and acid amount was found in different metal-doped SBA-15 and Ti/SBA-15 with different Ti loadings. 4Ti/SBA-15 has rich Lewis acid sites and exhibits the most excellent glycolysis ability; the BHET yield reaches 87.2% within 45 min at 190 °C (isolated yield 73%). The exception is 4Nb/SBA-15, although it has relatively abundant acid sites, its ability to depolymerize PET is still poor. 4Ti/SBA-15 was further applied to depolymerize common real PET plastics, such as dyed polyester fabric, and PET bottles, and the yield of BHET is still maintained at about 85%. Importantly, this catalyst also has high stability, and the yield of BHET (∼80%) does not decrease significantly during four cycles. In addition, in the scale-up experiment, the PET conversion is nearly 100% and the isolated BHET yield reaches 74%.
Catalysts with different metal centers were prepared by the incipient wetness method. Firstly, SBA-15 was added to a certain concentration of metal precursor solution, and then xM/SBA-15 was obtained after drying and calcination in turn, where x represented the mass fraction of the corresponding metal M. The calcination conditions: air atmosphere, 500 °C, heating rate 3 °C min−1, 4 h.
NH3-temperature-programmed desorption (NH3-TPD) were carried out on a PX200 desorption apparatus. Typically, 50 mg catalyst was pretreated at 500 °C in Ar (50 mL min−1) for 1 h and then cooled to 90 °C. The atmosphere was switched to 5% NH3/Ar (50 mL min−1), and the catalyst absorbed NH3 for 40 min at 90 °C. After the adsorption was completed, the atmosphere was switched back to Ar (50 mL min−1) and kept at 90 °C for 1 h to purge the residual NH3 in the quartz tube. Finally, the temperature in the quartz tube increased from 90 °C to 500 °C at a heating rate of 10 °C min−1, and the NH3 signal was detected by a TCD detector.
Fourier Transform Infrared (FT-IR) spectra for pyridine adsorption were recorded on a Nicolet NEXUS 670 FT-IR spectrometer. All the catalysts were pretreated at 400 °C under vacuum for 1 h to remove physically adsorbed water and impurities on the catalyst surfaces. Next, the cell was cooled to 200 °C, and the FT-IR spectrum was recorded as the background. The temperature in the cell was continued to cool to room temperature, and pyridine vapor was then introduced into the cell at room temperature until equilibrium was reached. Subsequently, the temperature was increased to 200 °C and held for 30 min to achieve equilibrium between pyridine adsorption–desorption. After the equilibrium was reached, a second FT-IR spectrum was obtained. The spectra presented were obtained by subtracting the spectra recorded before and after pyridine adsorption.
The DRIFTS of acetone adsorption was recorded on a NICOLET iS50 FT-IR spectrometer equipped with an MCT/A detector. Firstly, the catalysts were pretreated in situ in the cell in Ar at 400 °C for 1 h, and the background spectra were recorded at 190 °C and 35 °C, respectively. Then, acetone with Ar was bubbled into the in situ cell for 40 min. Next, the cell was purged with Ar at 35 °C for 30 min, and the adsorption spectra of acetone were recorded. Finally, the temperature was increased to 190 °C, and the adsorption spectra were recorded. In addition, the DRIFTS of acetone adsorption involving the pretreatment of methanol were also recorded on the above instrument. The difference is that methanol needed to be bubbled into the in situ cell in advance before acetone.
(1) |
(2) |
(3) |
The catalyst reusability experiment was carried out in the following steps. When the reaction was finished, acetonitrile was added to dissolve the BHET as well as the oligomers. The catalyst was then separated by centrifugation. The obtained catalyst was used directly in the next PET glycolysis reaction without calcination.
BHET was isolated by the method of József Kupai and co-workers.16 Firstly, acetonitrile was removed from the reaction solution by distillation under reduced pressure, and then deionized water was added to the remaining solution. At this time, flocculation precipitates appeared in the solution, which was the BHET dimer. After BHET dimer was removed by filtration, the resulting filtrate was distilled to about 4 mL under reduced pressure. At last, the pure BHET crystal was obtained by storing the resulting solution in a refrigerator at 4 °C for 12 h. After filtration, washing, drying, and weighing, the isolated BHET yield can be calculated.
Next, we would like to find the differences of various M/SBA-15 catalysts and to correlate their properties with the catalytic performance. Obviously, here, the acid sites in these four investigated catalysts are the main catalytic active centers in PET glycolysis. Therefore, the NH3-TPD was used to explore the acid properties in different metal-doped SBA-15 catalysts. As shown in Fig. 2a and Table S1,† the amount of acid sites of 4Sn/SBA-15 (0.04 mmol g−1) is much less than that of the other three catalysts, which explains its lowest PET glycolysis activity. Meanwhile, 4Ti/SBA-15 and 4Zn/SBA-15 have abundant acid sites (0.24 mmol g−1, 0.35 mmol g−1, respectively) and thus show excellent glycolysis activity. In addition, it should be noted that although 4Zn/SBA-15 has the largest amount of acid sites, its catalytic activity is slightly lower than 4Ti/SBA-15, which may be caused by its lower specific surface area (Table S1†). The same rule appears in Ti/SBA-15 catalysts with different Ti loadings. With the increase in Ti loading, the amount of acid sites in Ti/SBA-15 catalysts increases and the BHET yield then gradually increases (Fig. 2b and Table S2†). An intriguing question is why 4Nb/SBA-15, which has similar acid amount to 4Ti/SBA-15 (0.22 mmol g−1), exhibits poor PET glycolytic ability. Therefore, 4Ti/SBA-15 and 4Nb/SBA-15 were characterized in more detail to understand the reason for the difference in the catalytic activity between them.
Fig. 2 (a) NH3-TPD curves of various M/SBA-15 catalysts and (b) Ti/SBA-15 with different Ti loadings. |
Fig. 3a and b show the small- and wide-angle XRD patterns of 4Ti/SBA-15 and 4Nb/SBA-15. In the small-angle region, 4Ti/SBA-15 and 4Nb/SBA-15 have three distinct diffraction peaks at 2θ = 0.8–2.0°, corresponding to the (100), (110), and (200) planes of the well-ordered P6mm structure in the sequence from left to right.34 It indicates that the doping of Ti or Nb has little influence on pore ordering. In the wide-angle region, 4Ti/SBA-15 or 4Nb/SBA-15 has a characteristic band at 2θ = 15–35°, attributed to the amorphous silica framework. Fig. 3c shows the N2 adsorption–desorption isotherms of 4Ti/SBA-15 and 4Nb/SBA-15. The isotherms of both samples are type IV possessing a H1 hysteresis loop, indicating that both samples are mesoporous materials with well-ordered, regular, cylindrical pores. In addition, the specific surface areas of 4Ti/SBA-15 and 4Nb/SBA-15 are 835 m2 g−1 and 957 m2 g−1, respectively (Table S1†). HAADF-STEM images (Fig. 3d and g) show that the long-range ordered structure and mesoporous structure of SBA-15 are maintained when Ti or Nb species is doped, which further confirms the characterization results of XRD and N2 adsorption–desorption. The EDS mappings (Fig. 3f and i) reveal that Ti or Nb species is uniformly distributed on the siliceous SBA-15. The acid types of 4Ti/SBA-15 and 4Nb/SBA-15 were further investigated by FT-IR of pyridine adsorption. As shown in Fig. 4, the sharp 1450 cm−1 band is seen for both catalysts; meanwhile, no signal is observed at 1540 cm−1. It is evident that the 4Ti/SBA-15 and 4Nb/SBA-15 catalysts have a majority of Lewis acid sites, in line with the results reported in the literature.26,27
The above characterizations show that 4Ti/SBA-15 and 4Nb/SBA-15 have very similar textural properties and acidic properties. Hence, it is puzzling why 4Nb/SBA-15 acts poorly in PET glycolysis. It is accepted that the activation of CO bonds in PET chains is the key step in the reaction of PET glycolysis catalyzed by Lewis acid catalysts. But, on the other hand, if EG, another reactant and also a solvent, is adsorbed too strong on the catalyst surface, it would poison the Lewis acid sites, which was found in a methanol-poisoned system before.35 Therefore, we further verified the effect of alcohols on the adsorption–activation of CO bond over two catalysts. Acetone as a simple CO bond-containing molecule was used to perform adsorption–desorption DRIFTS experiments on 4Ti/SBA-15, 4Ti/SBA-15-methanol (pre-adsorption of methanol for 40 min), and 4Nb/SBA-15, 4Nb/SBA-15-methanol. For gas-phase acetone, the stretching vibration peak of the CO bond is at 1730 cm−1 band.35,36 A significant red shift is observed for both 4Ti/SBA-15 and 4Nb/SBA-15, which is located at 1710 cm−1 and 1712 cm−1, respectively. When these two catalysts are pretreated with methanol, the CO bond signal on 4Ti/SBA-15 hardly changes, while the peak area of the CO bond signal on 4Nb/SBA-15 decreases obviously, despite the red shift being preserved, suggesting that part of the active sites are poisoned by methanol. When the temperature increases to 190 °C (reaction temperature), the stretching vibration peaks of the CO bond on these two catalysts are both blue shifted after methanol pretreatment, and the blue shift is more intense on 4Nb/SBA-15 (Fig. 5b). It indicates that the active sites of 4Ti/SBA-15 exhibit weak adsorption of methanol and therefore have little effect on the subsequent activation of the acetone CO bond. For 4Nb/SBA-15, the strong oxygen affinity leads to the strong adsorption of methanol, which poisons the active sites35 and leads to weak CO bond activation. In view of this, we reasonably believe that 4Ti/SBA-15, which is dominated by Lewis acid sites, can activate CO bonds in PET chains under the reaction conditions, and PET can be effectively depolymerized. However, 4Nb/SBA-15 will be poisoned by EG due to its strong oxygen affinity, resulting in the inactivation of CO bonds in the PET chains. Thus, PET glycolysis cannot be carried out efficiently.
Fig. 5 Acetone DRIFTS spectra of 4Nb/SBA-15, 4Nb/SBA-15 pretreated with methanol, 4Ti/SBA-15, and 4Ti/SBA-15 pretreated with methanol recorded at (a) 35 °C and (b) 190 °C, respectively. |
Under the optimal conditions of PET glycolysis, the pure BHET crystal is obtained by special separation methods,16 and the isolated yield reaches 73%. The purity of the thus-obtained BHET crystal was further verify by 13C NMR spectroscopy and DSC (Fig. 7). The signals at 59.45, 67.50, 129.99, 134.22, and 166.65 ppm in Fig. 7a are attributable to the BHET monomer reported in the literature.14,37 The signal at about 40 ppm is attributable to the solvent, DMSO-d6. Furthermore, the signal at about 63.0 ppm corresponding to the BHET dimer is not observed in this spectrum. The DSC curve in Fig. 7b shows that only an endothermic peak corresponding to the melting point of BHET crystals is observed at 113 °C, while the melting point peak of the BHET dimer is not observed at 170 °C,38 further confirming the results of the 13C NMP spectrum, i.e., the BHET is pure after separation. Moreover, when maintaining the same ratio of catalyst and EG to PET but increasing the amount of PET to five times, the PET can still be fully depolymerized, with the BHET yield reaching 86.3% (isolated yield 74%, Fig. 7c). Finally, the 4Ti/SBA-15 catalyst was used for common real PET plastics, such as dyed polyester fabric and PET bottles, and it was found that the yield of BHET still reached about 85% (Fig. 7d and e). It is important to note that the pigments in dyed polyester fabric do not affect the PET glycolysis process but will affect the purity of the obtained BHET (pigments will residue in BHET, Fig. S3†).
Fig. 8 shows the reusability of 4Ti/SBA-15 in the glycolysis of PET. In four cycles, 4Ti/SBA-15 can efficiently catalyze PET glycolysis, and the yield of BHET is maintained at about 80%. The slight decrease in BHET yield may be due to the inevitable physical losses during the recovery of the catalyst. Inspiringly, 4Ti/SBA-15 still shows excellent catalytic activity and stability in a mixed plastic (PE + PS + PET, Fig. S4†). Fig. S5† shows the small-angle and wide-angle XRD patterns of 4Ti/SBA-15 after the reaction. In the small-angle and wide-angle range, the diffraction peak of 4Ti/SBA-15 is almost unchanged compared with the fresh catalyst, indicating that it has stable structural properties.
Fig. 8 The reusability of 4Ti/SBA-15 in the glycolysis of PET. Reaction conditions: 0.4 g PET, 2.8 g EG, 30 mg 4Ti/SBA-15, 1 MPa N2, 196 °C, 45 min. |
Fig. 9 (a) DR UV-vis spectra of Ti/SBA-15 with different Ti loadings. (b) The linear relationship between the absolute content of tetrahedral Ti4+ species and the BHET yield. |
The reaction process of PET glycolysis is proposed as follows (Scheme 2). First, the CO bond in the PET chain is activated by the tetrahedral Ti4+ species, causing the electrons in the CO bond to deviate toward the oxygen atom. And then, the carbon atom is vulnerable to attack by the free lone pairs of electrons on the oxygen of EG. Finally, the C–O bond in the PET chain is broken and a new C–O bond with EG is formed. Repeating in this way, the polymer PET was eventually depolymerized into BHET monomer.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cy01127e |
This journal is © The Royal Society of Chemistry 2023 |