The sol-gel method has been adopted successfully for the synthesis of Ag-doped ZnO nanoparticles. ZnO doped with 0.5, 1.0, 2.0, and 3.0 mol% was synthesised using the reported procedure (Vinitha et al. 2022). Depending on the Ag loading, Ag-doped ZnO was obtained as a light grey to grey powder. The crystallinity, crystallite size, optical bandgap, particle size, and elemental composition are determined using XRD, UV-Vis, SEM-EDX, and TEM-SAED techniques.
Diffraction and crystallite size analysis
Figure 1 shows the X-ray diffraction pattern of pure and Ag-doped ZnO nanoparticles recorded between 20 and 80˚. Strong and sharp peaks indicate the ordered wurtzite structure of ZnO nanoparticles. In addition, diffraction peaks related to AgO appeared to confirm the doping of Ag II (Fig. 1a). Diffraction angles shown at 31.70˚, 34.50˚, 36.30˚, 47.75˚, 56.65˚, 63.10˚, 66.35˚, 66.50˚, 67.90˚ and 69.95˚ accounted for planes (100), (002), (101), (102), (110), (103), (200), (112), (201) and (004) respectively according to the ZnO standard (JCPDS: 036-1451). The diffraction pattern of Ag-doped ZnO is depicted in Figs. 1a and 1b. The characteristic peaks observed at 38.15˚, 44.30˚ and 64.60 reflect the hkl values of planes (111), (200), and (220) corresponding to crystalline AgO-doped ZnO nanoparticles.
Based on Fig. 1a, a gradual increment in the intensity of AgO diffraction peaks indicated an increasing doping level of Ag (II) ions. Three major planes (100), (002) and (101) have not shown any change in diffraction angles (Fig. 1b). Therefore, the diffraction pattern of ZnO is not altered by Ag dopant in the ZnO framework (Fig. 1b). Based on the diffraction analysis, the crystallite size of Ag doped ZnO was calculated using the Debye-Scherrer equation (Eq. 1).
\(D=\frac{\text{K}{\lambda }}{\beta \text{cos}\theta }\) ……………… (Eq. 1)
D is crystallite size in nm, K is structural constant (value taken as 0.9), l is the wave length of X-ray source used for the diffraction studies (l = 1.5406 Å), is diffraction angles (in radians) and \(\beta\) is full width half maximum (FWHM) of diffraction peaks. Alteration in crystallite size, cell parameters and structural strain due to Ag doping calculated from Williamson-Hall method (Eq. 2).
\(\beta \text{cos}\theta =\frac{\text{K}{\lambda }}{D}+4ϵsin\theta\) ………….. (Eq. 2)
D is crystallite size (in nm), is diffraction angles (in radians) and e is lattice strain. Eq. 2 is a linear equation and Sin\(\theta\) is plotted against \(\beta\)Cos\(\theta\). From the plot, crystallite size D is calculated by rearranging intercept to D = K\({\lambda }\)/c and lattice strain (e) obtained from slope divided by 4.
The crystallite size calculated using the Debye-Scherrer equation revealed that crystallite size increased with increasing Ag dopant concentration (Table 2). In contrast, the W-H calculation method showed decreasing the crystallite size while increasing the concentration of Ag dopant, and the results are consistent with the earlier reports (Nigussie et al. 2018; Thakur et al. 2020; Rahman et al. 2019). Furthermore, the lattice strain (ε) increased as the Ag(I) concentration increased from 0.5 to 3.0 mol%. The decrease in crystallite size and increase in lattice strain indicates the incorporation of a large ionic radius (1.55 A°) in Ag(I) as opposed to a smaller ionic radius (0.74 A°) in Zn(II). The particle size of Ag-doped ZnO was determined from SEM images of respective nanoparticles using the software imageJ (version 1.53) to support the crystallite size (Fig. 4). The agglomeration of smaller Ag-doped ZnO results in the formation of nanoparticles with particle sizes ranging from 118 to 135 nm, with a gradual increase in particle size, observed as the concentration of Ag(I) doping increases.
Table 2
Lattice parameter, crystallite size and lattice strain of ZnO and Ag doped ZnO
Material
|
Lattice parameters
|
volume (nm3)
|
Crystallite size (nm) (Scherrer equation)
|
Crystallite size (nm) (W-H equation)
|
Lattice strain \({\epsilon }\) X 10− 3
|
SEM
Particle size (nm)
|
a (Å)
|
c (Å)
|
c/a
|
ZnO
|
1.69
|
5.19
|
3.05
|
12.98
|
37.65
|
80.40
|
1.40
|
156.92
|
0.5mol% Ag-ZnO
|
1.69
|
5.19
|
3.06
|
12.96
|
50.48
|
60.34
|
0.275
|
118.39
|
1.0mol% Ag-ZnO
|
1.90
|
5.20
|
2.74
|
16.18
|
49.54
|
51.72
|
0.375
|
118.01
|
2.0mol% Ag-ZnO
|
1.90
|
5.19
|
2.74
|
16.17
|
57.42
|
43.88
|
0.475
|
125.47
|
3.0mol% Ag-ZnO
|
1.93
|
5.19
|
2.68
|
16.82
|
61.90
|
36.24
|
0.525
|
134.48
|
UV-Vis studies and Optical bandgap calculation
Figure 2a shows the UV-Vis absorption spectra of pure and Ag-doped ZnO nanoparticles, recorded in the range of 200–800 nm. The outer shell electrons of ZnO and Ag2O are excited to a higher energy state by absorbing light to give an absorption peak. The absorption spectra give valuable information on bandgap alteration by the dopants in any kind of semiconducting metal oxides, such as TiO2 and ZnO. The optical bandgap of Ag-doped ZnO NPs is calculated from the onset of the absorption peak in the higher wavelength region. Ag/ZnO NPs showed an absorption maximum in the range of 350–360 nm, and a change in absorption maxima was not observed due to doping. However, 2 and 3 mol% Ag(I) doped ZnO showed second absorption maxima at 554 and 556 nm, respectively, due to the presence of Ag2O, and visually, these two samples appeared pale grey and dark grey, respectively.
Table 3
UV-Vis absorption maxima, optical band gap, FT-IR and elemental composition
Material
|
Absorption maxima
(λ, nm)
|
Optical band gap (eV)
|
Wave number (cm− 1)
|
FT-IR assignments
|
Weight (%)
|
Atomic (%)
|
λ1
|
λ2
|
ZnO
|
350
|
-
|
3.26
|
440
|
Zn-O
|
-
|
-
|
0.5 mol% Ag-ZnO
|
351
|
-
|
3.21
|
440
500
990–1010
1090–1110
|
Zn-O
Ag-O
Zn-O sym
Zn-O sym
|
Zn – 80.24
O – 19.92
Ag – 0.63
|
Zn − 50.54
O – 49.24
Ag – 0.24
|
1.0 mol% Ag-ZnO
|
350
|
-
|
3.21
|
|
|
Zn – 75.20
O – 24.80
Ag – 0.65
|
Zn – 42.20
O – 57.40
Ag – 0.25
|
2.0 mol% Ag-ZnO
|
350
|
554
|
3.19, 1.95
|
440
500
990–1010
1090–1110
|
Zn-O
Ag-O
Zn-O sym
Zn-O sym
|
Zn – 79.20
O – 19.79
Ag – 1.01
|
Zn – 49.29
O – 50.33
Ag – 0.38
|
3.0 mol% Ag-ZnO
|
350
|
556
|
3.21
|
440
500
990–1010
1090–1110
|
Zn-O
Ag-O
Zn-O sym
Zn-O sym
|
Zn – 79.71
O – 19.65
Ag – 1.95
|
Zn – 49.69
O – 50.07
Ag – 0.89
|
The bandgap values were obtained from Tauc’s plot by plotting (αhϑ)−1/2cm−1/2 against photon energy in eV. Figures 2b and 2c show Tauc’s plot of pure and Ag doped ZnO NPs. The onset of Tauc's plot curves directly gives an optical bandgap in eV. Change in optical band gap is observed due to doping (Table 2). According to Tauc's plot, pure ZnO had an optical band of 3.26 eV, while Ag(I)-doped ZnO had optical bands of 3.21 and 3.19 eV for 2 and 3 mol% doped ZnO NPs, respectively. A decrease in optical bandgap was observed, and in addition, another bandgap was observed at 1.95 eV due to 2 and 3 mol% Ag doping (Bagha et al. 2021; Yousefi et al. 2012).
FT-IR analysis
FTIR spectroscopy is used to determine the molecular structure and various functional groups present in the pure and Ag-doped ZnO NPs. Figure 3a shows the FTIR spectra of pure and 0.5, 1.0, 2.0, and 3.0 mol% of Ag-doped ZnO nanoparticles synthesised by the sol-gel method and calcined at 500 ˚C. Figure 3b depicts the expanded position of the first absorption band. FT-IR spectra are recorded in the range from 400 to 4000 cm− 1. The position and number of absorption bands depend on the crystal morphology, crystal structure, and chemical composition. All the samples exhibited broad and sharp absorption bands in the range of 420 to 540 cm-1 attributed to the metal oxygen stretching vibration modes and clearly confirmed the formation of Zn-O and Ag-O stretching bond. This indicates the successful incorporation of Ag ions into the crystal lattice of ZnO. Furthermore, the two broad peaks with approximate intensities ranging from 980 to 1140 cm− 1 are associated with symmetric and asymmetric stretching vibrational states, respectively, whereas the Zn-O mode, which extends peak intensities at 510 cm− 1, is more common in undoped ZnO wurtzite (Rajendran, 2020; Sagadevan, 2017; Ramesan, 2018).
SEM morphological analysis
The morphological structures of the prepared nanoparticles were examined by SEM analysis, as shown in Figs. 4a and 4b for 0.5, 1.0, 2.0, and 3.0 mol% Ag doped ZnO, respectively. The SEM-EDX images of pure and different mol% of Ag-doped ZnO NPs composite at two different magnifications. EDX analysis is shown in Fig. 4c Doped and undoped ZnO NPS were magnified at the scale of 1µm and 500 nm is shown in 4a and 4b SEM. The images clearly show the spherically shaped Ag-doped ZnO NPs at all magnifications (Figs. 4a (1) and 4b (1)). The microstructure revealed a uniform and compact structure, interconnected by grains of 100–120 nm size on average. Also, the morphology of the particles is approximately “spherical” in shape. The morphology of integrated Ag-ZnO NP was not affected by the incorporation of Ag ions into the ZnO lattice.
Energy-dispersive X-ray spectroscopy (EDX) (Fig. 4c) is used to determine the basic elemental composition of the synthesised Ag-doped ZnO nanoparticles. The EDX spectra of 0.5, 1, 2, and 3 mol% Ag-doped ZnO NPs are shown in Fig. 4c. EDX analysis confirms the presence of the dopant Ag in the Ag-ZnO nanoparticles. In the observed EDX spectra, well-defined peaks for zinc (Zn), oxygen (O), and silver (Ag) are found. No impurity or peak other than Zn, Ag, and O atoms were observed in the EDX spectrum, which revealed that the integrated NPs were composed of Zn, Ag, and O. The observed EDX results are well-matched with the diffraction pattern of XRD. The composition of 0.5 mol% Ag doped ZnO contains 0.65 wt% and 0.25 atomic% of Ag, 2 mol% Ag doped ZnO contains 1.01 wt% and 0.38 atomic% of Ag, and 3 mol% Ag doped ZnO contains 0.63 wt% and 0.24 atomic% of Ag, which is consistent with ZnO doped with different mol% Ag. (Dhatshanamurthi, 2016; Gnanaprakasam, 2016).
TEM
High-resolution TEM (HRTEM) was used to examine the detailed morphological and structural properties and composition of as prepared 2 mol% Ag-doped ZnO NPs (Kumar, 2015). Figure 5a and b show the TEM images of 2 mol% of Ag-doped ZnO nanoparticles reached at scales of 50 and 100 nm, respectively. Figure 5c and d show the SAED and EDX spectra and elemental composition of 2 mol % of Ag-ZnO nanoparticles. The TEM images, it is clearly shown that the Ag-doped ZnO has a hexagonal structure and a spherical shape. This HRTEM result agrees with the XRD result, indicating that Ag-doped nanoparticles have a higher crystalline nature and that the Ag nanoparticles are uniformly distributed over the ZnO lattice. Both images show the typical hexagonal wurtzite structure that is consistent with their XRD results. Figure 5c shows an image of selected-area electron diffraction (SAED) confirming the patterns of Ag-doped atoms with circular rings as planes (111), (200), and (220), corresponding to polycrystalline in nature (Ali, 2018). The elemental mapping and composition of Ag-doped ZnO NPs are determined using EDX spectra. The EDS spectra of 2 mol % Ag-ZnO NPs are shown in Fig. 5e. The peaks of elements Zn, O, and Ag have been identified in the EDX spectra. The EDX spectra reveal that Ag-doped ZnO NPs are composed mainly of Zn, O, and Ag elements. The composition of 2 mol% Ag doped ZnO is 9.96 wt% and 4.75 atomic% Ag, which is consistent with ZnO doped with 2 mol% Ag (Tsai, 2018; Suresh Kumar, 2015; Ramesh Kumar, 2015).
Depolymerization Of Polyester Threads Using Aminolysis And Glycolysis
Figure 6 summarised glycolytic and aminolytic depolymerisation of red, blue and green polyester threads using 2 mol% Ag doped ZnO NPs and compared them with pure ZnO NPs. All the reactions carried out in the presence of 5wt% of catalyst to PES and, ethanol amine or diethanol amine was used as aminolyting agents, similarly, ethylene glycol(EG) and diethylene glycol(DEG) (PES: EG, 1:20) used as a glycolytic agent (Scheme 1). The aminolysis of PES threads using ethanol amine catalysed by ZnO and 2 mol% Ag doped ZnO afforded more than 94% of bis(2-hydroxy ethyl) terephthalamide (BHETA). The nucleophilic centers in diethanolamine (DEA) and ethanolamine (EA) are hydroxyl and amine. Because nitrogen is a good nucleophile than oxygen, the amine group of DEA and EA attacks the ester linkage of the PET during aminolytic depolymerization.
Optimisation Of Reaction Parameters On The Aminolysis Reaction Of PES Textile Waste
Effect of Ag doping
To study the effect of different mol% of Ag-doped ZnO nanoparticle catalyst on BHETA yield, we used ethanolamine as the aminolytic agent and a constant ratio of 1:20 PET to EA at 180 W under microwave irradiation for 30 h, as shown in Fig. 7a. It can be observed that increasing the mol % ratio of Ag doping waste results in a significant increase in BHETA yield. Among all the catalysts, 2 mol% of Ag-doped ZnO NPs produced the highest BHETA yield, which was greater than the ZnO-catalyzed reaction yield.
Effect of microwave power
Figure 7b showed the PES textile waste decomposition was carried out at different MW of radiation power.The reaction time was kept constant for 30 min with irradiation at 50, 100, 150, 180, 200, and 220 W. Even after 30 minutes of irradiation, the PET degradation is only 50% for the lowest power used (50 W), and when 180 W is used, almost complete dissolution occurs within 2 minutes. After 20 minutes of irradiation at 75 W, complete PET depolymerization occurs. The results indicate an increase in the yield of BHETA, increasing microwave power from 150 to 200 W. This indicates that 180 W of microwave power is optimal for irradiation, as demonstrated by the subsequent reactions.
Effect of irradiation time
To investigate the effect of reaction time on BHETA yield, we performed aminolysis of PES textile waste at 180 W with ethanolamine as the aminolytic agent. The resulting yield is shown in Fig. 7c. The complete depolymerization of PES textile waste under 180 W microwave irradiation was observed in the reaction time range of 2 to 15 min. In the time range of 10 to 15 minutes, an increase in PES conversion from 60 to 90 percent is observed. PES textile waste was completely (100%) converted after 15 min, and the BHETA yield was 96%. Finally, the results indicated that the yield of BHETA and the conversion of PES textile waste increased with increasing irradiation time. This showed that time has been found to be a major factor in the depolymerization of PES textile wastes. In order to find the minimum reaction time required for complete depolymerization of PES, 100 mg of 2 mol% Ag-ZnO and a 1:20 PES:EA ratio were used for the aminolysis reaction.
Effect of temperature
Figure 7d showed the effect of reaction temperature on the depolymerization of PES textile waste under microwave irradiation, temperature is a crucial factor in aminolysis experiments. Temperatures of 100, 120, 150, 180, and 200 ˚C were used at 10- to 15-minute intervals, along with 100 mg of 2% Ag doped ZnO catalyst, 20 ml of ethanolamine, and 500 mg of PES textile wastes. As predicted, increasing the reaction temperature causes more PES textile waste depolymerization and BHETA formation. Even after 60 minutes of irradiation at low temperatures of 100 and 120 ˚C, PES textile waste decomposition was only 55 and 70 percent, respectively. When the temperature is raised to 150–180 ˚C, the dissolution of PES textile waste takes between 5 and 15 minutes to achieve more than 90% depolymerization. It can be assumed that using microwave irradiation significantly reduces the time required for PES textile waste conversion and increases the yield of BHETA.
Recycling of catalyst
2 mol% Ag-doped ZnO nanoparticle catalyst was investigated for catalyst recycling studies shown in Fig. 7e. For economic and environmental reasons, most chemists are focusing on determining the viability of catalyst recycling. Therefore, the utilisation and recovery of the catalyst under optimised conditions were explored for further amionolysis of PES textile waste. Following filtration, the catalyst was washed with methanol and calcined at 550 ˚C for 3 hours. In this method, the Lewis acid catalyst Ag-doped ZnO can be recycled up to six times with negligible activity loss. However, the decrease in BHETA yield observed is due to a stronger interaction of ZnO Lewis acidic sites with ethanol amine, which in turn affects the structure of ZnO.
Aminolysis of PES threads using ethanol amine (EA) and diethanol amine (DEA)
Figure 8a shows PES textile waste recycling through the aminolysis process. PES textile waste included red, blue, and green threads, with EA and DEA used for aminolysis reactions. The reaction of each EA and DEA with PES results in the splitting of the ester bonds of PES and the formation of the amides BHETA and BDHETA, respectively. The amide is the main reaction product due to the higher reactivity of primary amines in ester bond cleavage. More than 90% of bis (2-hydroxyethyl) terephthalamide (BHETA) was obtained from aminolysis of PES threads using ethanol amine catalysed by ZnO and 2 mol% Ag doped ZnO. At 180°C and a PET/EA ratio of 1:20, 2 mol% Ag-ZnO NPs were obtained in 15 minutes, with the highest PES thread conversion and BHETA yields of 95, 95, and 96% for red, blue, and green threads, respectively. Furthermore, the DEA's requirements for producing BDHETA were met at a rate of 80–90%.EA is found to be a good aminolyzing agent for PES textile waste degradation because it produces a higher BHETA yield than DEA.
Glycolysis of PES threads using ethylene glycol (EG) and diethylene glycol (DEG)
Several researchers are becoming more interested in PES textile waste recycling via microwave irradiation. The microwave recycling process enables relatives to take a short reaction time while using much more energy for heating than the traditional heating process. Even so, the benefits of microwave irradiation of PES textile waste via the glycolysis process are still prohibitively expensive due to energy consumption. Glycolysis is the most cost-effective and commercially viable method of chemically recycling PES textile waste. Glycolysis is a type of solvolysis reaction in which glycol is used as the reactive solvent. The glycolysis reaction is chemically defined as the "molecular degradation of PET polymer by glycols in the presence of trans-esterification catalysts, where the ester linkages break and are replaced with hydroxyl terminals."
Figure 8b depicts a comparison of the glycols performance of pure and Ag-doped ZnO NPs catalysts in the glycolysis reaction of PES textile waste. The reactor was placed in a microwave, and the reaction microwave power was maintained at 300 watts for 30 minutes. PES/EG and PES/DEG reaction temperatures were in the 180–200 ˚C range at 300 watts of power. Glycolysis commonly uses EG to produce BHET and DEG to produce BHEOET. The polyester threads (red, green, and blue) reacted with EG when Ag-ZnO NPs were used as a catalyst, and the BHET yield increased to 90%, compared to slightly higher yields from pure ZnO. More than 80% of the BHEOET produced was made possible by polyester threads treated with DEG. The reaction resulted in 100% depolymerization of PES textile waste, 80 to 90% yields of BHET and BHEOET, and more than 10% oligomer.
Spectral Characterisation Of BHETA
The complete depolymerization of PET wastes generates BHETA and BHET as a single product. However, there could be a possibility of forming dimer and oligomer products due to incomplete depolymerization. Thus, spectral characterization techniques such as MS and FT-IR analysis have been carried out. To confirm the purity and structural confirmation of the depolymerized products, BHET and BHETA, structural characterization was performed. FT-IR of BHETA and BHET is given in Figs. 9 and 10, respectively (the full spectrum is given in the supplementary information), as is their chemical structure. The 1HNMR-depolymerized PES waste showed up in Figs. 11 and 13. The ESI-MS spectra of BHETA and BHET are given in Figs. 14 and 15, respectively.
FT-IR spectrum of BHETA and BHET
Figures 9a, b, and c show the FTIR spectra of the aminolyzed product of BHETA, which were recorded from 400 to 4000 cm− 1. The presence of NH and hydroxyl groups was confirmed by a broad absorption band around 3354 cm− 1 and 3280 cm− 1. Broad bands at 3089 and 2958 cm− 1 were assigned to the aromatic C-H stretching vibrations and aliphatic CH2 vibrations. The presence of amide linkages (-CONH-) in the structure was confirmed by a sharp peak at 1616 cm− 1. The presence of aromatic C = C stretching vibrations is responsible for the peak at 1548 cm− 1. The peak at 1047 cm− 1 is due to aliphatic C–H stretching vibrations. The presence of BHETA was confirmed by the presence of aromatic and aliphatic -OH and -NH stretching vibrations, as well as an amide carbonyl vibration (Jamdar, 2017; Musale, 2016).
As shown in Figs. 10a, b, and C, Ag-doped ZnO NPs catalysed the glycolysis reaction of PES textile waste of red, blue, and green threads, which are BHET products.The main product's FTIR spectrum agrees well with those described in the literature for BHET. The presence of hydroxyl groups is indicated by the peak at 3437.15 cm-1 (OH). The bands at 2960 to 2879 cm-1 correspond to the stretching vibrations of CH2 domains, respectively. The peaks at C = O, C-O ester bond asymmetric vibration, and C-O ester bond symmetric vibration are assigned to the bands at 1705, 1261, and 1066 cm-1, respectively. The weak band at 1408 cm-1 is very typical of aromatic C = C stretching. The presence of aromatic residue in the glycolyzed products is strongly suggested by the bending frequencies at 721 and 489 cm-1. Furthermore, the sharpness of the FTIR peaks suggests that the BHET obtained from the depolymerization process is of high purity (Sert, 2019; Mendibura, 2021; Hoang, 2018).
As shown in Figs. 10a, b, and c, Ag-doped ZnO NPs catalysed the glycolysis reaction of PES textile waste of red, blue, and green threads, which are BHET products. The main product's FTIR spectrum agrees well with those described in the literature for BHET. The presence of hydroxyl groups is indicated by the peak at 3437.15 cm− 1 (OH). The bands at 2960 to 2879 cm− 1 correspond to the stretching vibrations of CH2 domains, respectively. The peaks at C = O, C-O ester bond asymmetric vibration, and C-O ester bond symmetric vibration are assigned to the bands at 1705, 1261, and 1066 cm− 1, respectively. The weak band at 1408 cm− 1 is very typical of aromatic C = C stretching. The presence of aromatic residue in the glycolyzed products is strongly suggested by the bending frequencies at 721 and 489 cm− 1. Furthermore, the sharpness of the FTIR peaks suggests that the BHET obtained from the depolymerization process is of high purity (Sert, 2019; Mendibura, 2021; Hoang, 2018).
NMR spectrum of BHETA and BHET
Proton 1H NMR is used to get a better insight into the chemical structure of BHET and BHETA. DMSO was used to obtain the 1H NMR spectrum of the depolymerized product at room temperature using a 500 MHz instrument. The NMR spectra of the final product of BHETA obtained from red (VSAR1), blue (VSAB2), and green (VSAG3) are shown in Figs. 11a, b, and c. The aminolyzed product, bis-(2-hydroxyethyl)terephthalamide (BHETA), is a symmetrical molecule. The 1H NMR spectra of BHETA revealed two NH protons bonded to amide carbonyl (C7 and C8) (-C = O) and methylene carbons (C9 and C10) (-CH2) as a triplet at 8.55–8.58 ppm with a J value of 5.5 Hz. The presence of NH protons shows depolymerization of PES and BHETA formation. The position and multiplicity of aromatic protons provide valuable information about the monomer, dimer, and oligomeric nature of depolymerized products. The skeleton of the BHETA molecule is symmetrical, as evidenced by the signal of aromatic protons attached to C1, C2, C4, and C5 appearing as a singlet at (Fig. 11a) 7.92 ppm and (Fig. 11b and c) 7.93 ppm, indicating that BHETA is the pure and single product formed, which indicates that depolymerization of PET wastes is complete. Furthermore, Fig. 11a showed up singlet at 4.77 ppm, and at 4.75–4.78 ppm with a J value of 5.5 Hz (Fig. 13b and c), two -OH protons attached to methylene carbons (C11 and C12) appeared as a triplet. Between 3.50 and 3.54 ppm (Fig. 11b and c), the protons of the methylene group (C11 and C12) attached to -OH appeared as a quartet with a J value of 6 Hz. All 1H NMR spectra show a peak corresponding to methylene protons (C9 and C10) attached to an amide NH group, which produces a triplet at 2.51 ppm and a J value of 6 Hz. The spectral values of the aminolyzed product of BHETA agree well with the spectral values published previously (-NH-CO-C6H4-CO-NH-) (Vinitha et al. 2022; Jeya et al. 2022a).
Figures 13a, b, and c depict the 1H NMR spectra of the depolymerized BHET final product [red (VSAR1), blue (VSAB2), and green (VSAG3)], with the corresponding peak assignments marked in the structural formula. The BHET is a symmetrical molecule, and the four aromatic (C1, C2, C4, and C5) hydrogens present in BHET have shown up as singlets at 8.14 ppm, confirming the formation of the glycolyzed product as the BHET monomer. The 1H NMR analysis revealed singlet aromatic protons, confirming the complete depolymerization of PES textile wastes. At the peak 4.97–5.01 ppm (Fig. 13a & c), the two hydrogens attached to carbon (C9 and C10) next to the ester carbonyl oxygen (-CH2-OC = O) (C7 and C8) appeared as a triplet, and the peak 4.97–5.01 ppm (Fig. 13b) appeared as a quartet with a J value of 5.5 Hz. Due to the exchangeable nature of its kind, some other four hydrogens (C11 and C12) connected to CH2 and OH appear as a triplet in the range of 4.31–4.35 ppm and are present in both the spectra and J value at 6 Hz. A hydroxyl group attached to CH2 appeared as a broad quartet in the 3.70–3.75 ppm range. At 2.50 ppm, the deuterated DMSO solvent appears as a singlet with a J value of 6 Hz. (Jeya et al. 2017, 2020, and 2022)
Mass Spectrum Of BHETA
Figure 14a, b and c represents the mass spectrum of the aminolyzed products, bis (hydroxy ethyl) terephthalamide obtained from red blue and green PES threads. BHETA's molecular formula is C12H16N2O4 and its molecular weight is 251.00 g/mol. The Red PES thread (VSAR1) (Fig. 14a) spectrum recorded using the electron spray ionisation method exhibited a protonated molecular ion peak at m/z 251.35 (M + H). The following peak at m/z 268.20 predicts that the Na+ Hydrated form of BHETA was formed as a result of electron spray ionisation. Then the mass spectrum of blue PES (VSAB2) (Fig. 14b) revealed the molecular ion peak at 251.05 (M + H) in the protonated form and the next peak at m/z 287.00 indicates the formation of dihydrated form. Finally, the last image of green-coloured PES (VSAG3) (Fig. 14c), which gives a peak at 250.10 (M + H) attributed to the main product of BHETA, one more peak at m/z 283.15 clearly shows that the Na+ form of BHETA due to electron spray ionization (Vinitha, 2022).
Mass Spectrum Of BHET
Figures 15a, b and c represent the mass spectrum of the glycolysed product, bis (2-hydroxyethylene) terephthalate obtained from red, blue and green PES wastes. BHET's molecular formula and the molecular weight is C12H14O6 and 254.00 g/mol, respectively. The mass spectra demonstrate that the main product was obtained with high purity, as no signals from other compounds were detected. The ESI-MS spectrum of color PES (VSGR1) (Fig. 15a) showed a molecular ion peak with m/z 254.25 (M + H) with 100% intensity, which corresponds to BHET ionised by Na + value is 294.40. A common feature in the figures of (15b and c) VSGB2 and VSGR3 showing mass spectrometry results is that the parent mass peak is located at 252.7 (M + H) In the case of molecular ions of Na+ form of BHET. Entirely these data indicate that highly pure BHET was obtained after the crystallization process in water was performed. Figure 15 (b) and (c) mass spectrum of BHET obtained from glycolytic depolymerisation of blue and green PES threads. Both the mass spectrum showed molecular ion peaks at m/z 252.75 and 252.70 respectively for BHET generated from blue and green PES wastes (Sce, 2019).