Next Article in Journal
Dilution-Induced Deposition of Concentrated Binary Mixtures of Cationic Polysaccharides and Surfactants
Next Article in Special Issue
Preparation and Characterization of Silica-Based Ionogel Electrolytes and Their Application in Solid-State Lithium Batteries
Previous Article in Journal
New Polyvinyl Alcohol/Succinoglycan-Based Hydrogels for pH-Responsive Drug Delivery
Previous Article in Special Issue
Synthesis and Characterization of Poly(DL-lactide) Containing Fluorene Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comprehensive Kinetic Study of PET Pyrolysis Using TGA

Chemical Engineering Department, King Faisal University, P.O. Box 380, Al-Ahsa 31982, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(14), 3010; https://doi.org/10.3390/polym15143010
Submission received: 27 May 2023 / Revised: 6 July 2023 / Accepted: 10 July 2023 / Published: 11 July 2023
(This article belongs to the Special Issue Advances in Polymerization and Kinetic Studies)

Abstract

:
The pyrolysis of polyethylene terephthalate (PET) is a well-known process for producing high fuel value. This paper aims to study the kinetics of PET pyrolysis reactions at 4 different heating rates (2, 5, 10, and 20 K min−1) using thermogravimetric analysis (TGA) data. TGA data show only one kinetic reaction within the temperature ranges of 650 to 750 K. Five different model-free models, namely, the Freidman (FR), Flynn–Wall–Qzawa (FWO), Kissinger–Akahira–Sunose (KAS), Starink (STK), and distributed activation energy model (DAEM), were fitted to the experimental data to obtain the activation energy (Ea) and the pre-exponential factor (A0) of the reaction kinetics. The Coats–Redfern (CR) model equation was fitted with the help of master plot (Criado’s) to identify the most convenient reaction mechanism for this system. Ea’s values were determined by the application of the five aforementioned models and were found to possess an average value of 212 kJ mol−1. The mechanism of PET pyrolysis reaction was best described by first-order reaction kinetics; this was confirmed by the compensation. Further thermodynamic parameter analysis indicated that the reaction was endothermic in nature.

1. Introduction

Plastic waste does not biodegrade easily and its steady increase in the last three decades constitutes a major concern for its environmental risk. Recycling of this great amount of waste has been investigated by many countries, scientists, and researchers, and pyrolysis degradation is one of the most promising solutions for the recovery of a high-value fuel product. TGA analysis was usually run at the first stage to collect the various kinetic parameters which are essential for reaction process design.
PET is recognized as one of the main six polymers in plastic waste, and about 7.6% of plastic waste is PET (Martín-Gullón et al. (2001) [1], Diaz-Silvarrey et al. (2018) [2]). PET is used in the manufacturing of fibers and films, especially for soft drink containers, because of its high stability characteristics (Moltó et al. (2007) [3], Çepelioğullar and Pütün (2013) [4]). Many researchers have reported the results of kinetics studies on PET pyrolysis reaction using TGA in the last twenty years. Some applied a simple “curve-fitting“ method to collect the kinetic parameters and hence reported Ea of 242 kJ mol−1 with one reaction order (Yang et al. (2001) [5]). Subsequently, the decomposition kinetics of PET polymer were studied by the isoconversional method and the Vyazovkin model-free method, and the activation energy was found to be strongly dependent on the conversion of reaction (Saha et al. (2006) [6]). Brems et al. (2011) [7] reported the average value of activation energy of 237 kJ mol−1 at wide ranges of heating rate (3–120 K min−1) for PET pyrolysis reaction, similar to the value obtained by FWO model-fitting calculation.
Çepelioğullar and Pütün (2013) [4] studied the pyrolysis of PET alone as a set of experiments. They used CR (model fitting) at a single heating rate of 10 K min−1 to calculate the kinetic parameters. They found 2 values of activation energy (347.4 kJ mol−1 for conversion of 0.4–81% and 172.6 kJ mol−1 for the conversion of 88–99%, respectively). They attributed this high value of activation energy to the complex compound with an aromatic ring structure. Diaz Silvarrey and Phan (2016) [8] carried out TGA analysis for the pyrolysis of PET at different heating rates (5, 10, 20, and 40 K min−1). They determined the value of Ea and the pre-exponential factor of 197.61 kJ mol−1 and 4.84 × 1014 s−1, respectively, using KAS and FR methods. Miandad et al. (2017) [9] found that the decomposition of PET occurred within 2 stages (peak 753 K and 923), results which contradicted the findings of Dimitrov et al. (2013) [10], and found that these differences were due to polymeric structure and the differences in degradation mechanisms (Chandrasekaran et al. (2015) [11]). Ganeshan et al. (2018) [12] applied the CR model to PET pyrolysis reaction at 3 different heating rates (15, 20, and 25 K min−1) with the 2-stage degradation profiles and found the Ea values to be between 133 and 251 kJ mol−1. They highlighted that the main initial decomposition in the range of temperature (600–740 K) occurred with more than 80% mass loss. They pointed out that the CR model analysis was not appropriate for evaluating the kinetic parameters (Osman et al. (2020) [13]). Das and Tiwari (2019) [14] used plastic Coca-Cola bottle waste as a PET sample for TGA pyrolysis at a wide range of heating rates (5, 10, 20, 40, and 50 K/min). They used only the advanced isoconversional (AIC) method and obtained activation energy values between 203 and 355 kJ mol−1. Further, they applied Criado’s master plot technique to establish F1 as a possible mechanism model for inert TGA data. Osman et al. (2020) [13] applied differential FWO, integral FR isoconversional, and kinetic modelling (ASTM-E698) to calculate the kinetic triplet for PET pyrolysis. They found different activation energy values with the above three methods with an error rate of less than 15% between the lowest and the highest value. Mishra et al. (2019) [15] used different isoconversional model-free methods (FR, KAS, FWO, STK, and CR) to obtain the kinetic parameters for PET pyrolysis. Average activation energy values for all these methods are presented in results and discussion section. They applied the CR method for a single heating rate in order to estimate the order of the reaction by adopting various models including the Avrami and diffusion model. They also calculated enthalpy, entropy, Gibbs free energy, and frequency factor as part of thermodynamic analysis using activation energy values obtained from the STK method. Chowdhury et al. (2023) [16] studied PET pyrolysis using FR (model-free), CR, and Arrhenius methods (model fitting) at 3 different heating rates (10, 20, and 30 K min−1). They calculated very different ranges of activation energy values (FR = 3.31 to 8.79 kJ mol−1, CR = 1.05 × 104 kJ mol−1, and Arrhenius = 1278.88 kJ mol−1). Moreover, they computed the thermodynamic parameters using the Arrhenius and CR models.
This current study aimed to obtain full information on PET pyrolysis using TGA experimental data. The kinetic parameters and the mechanism of the pyrolysis were collected comprehensively by five isoconversional methods (FR, FWO, KAS, STK, and DAEM) with two nonisoconversional methods (CR and Criado). In addition, thermodynamic parameters of the pyrolysis reaction process were calculated and interpreted from the model analysis.

2. Materials and Procedures

2.1. The Proximate and Ultimate Analyses

PET was collected from the same supplier (Recycled Plastic, Ipoh, Malaysia) as used during our previous experiment (Dubdub and Al-Yaari (2020) [17]). Two main characteristic (proximate and ultimate) analyses were performed to obtain the physiochemical properties of the sample, for which detailed results are presented in Table 1. The procedure is detailed elsewhere (Dubdub and Al-Yaari (2020) [17]).

2.2. TGA of PET

PET polymer samples which were collected from Recycled Plastic (Ipoh, Malaysia), were ground into powder before use in the TGA analysis. An amount of 10 mg of PET was used with 40 mL min−1 N2 gas flow inert atmosphere at 4 different heating rates (2, 5, 10, and 20 K min−1). Each test run of these four is indicated by PETx throughout the paper, and x number refers to the heating rate. Multiple heating rates were applied, following the recommendations of the ICTAC (Koga et al. (2023) [19]). Mishra et al. (2019) [15] noted that for accurate determination of kinetic parameters, a low heating rate (below 8 K min−1) should be used, and the ratio between the highest and lowest heating rate should be greater than 10 K min−1 (Osman et al. (2020) [13]).

2.3. Derivation of the Kinetic Equations

The derivation of PET pyrolysis reaction will be based on the following well-known Arrhenius equation:
d α d t = A 0 exp E a R T f α
where α is the reaction conversion, t is time, Ea is the activation energy, A0 is the frequency factor, R is the universal gas constant, and T is the absolute temperature (Dubdub and Al-Yaari (2020) [17]).
For nonisothermal test, β (heating rate) can be introduced in the above equation as:
β d α d T = A 0 exp E a R T f α
All five model-free methods can be derived from Equation (2) either integrally to obtain FR method or differentially to obtain FWO, KAS, STK, and DAEM with some assumptions for each method (Chowdhury et al. (2023) [16], Aboulkas et al. (2010) [20], Dubdub (2023) [21]). Table 2 presents these five isoconversional equations, and Table 3 shows the equations for CR and Criado. Criado attempted to verify the reduced theoretical curve (left side) and the experimental data (right side) in Equation (9). Therefore, a comparison between them will help us determine which kinetic model will describe the experimental reaction. Table 4 shows the common solid-state thermal reaction mechanisms f(α) and g(α) used in the CR and Criado method (Table 3).
By plotting ( ln β d α d T , ln β , ln β / T 2 ,   ln β / T 1.92 , ln β / T 2 ) against 1 T for FR, FWO, KAS, STK, and DAEM models, the value of Ea will be obtained from the slope of the line. The obtained values of Ea by these methods are independent of the reaction mechanism. CR model, expressed by Equation (9), was implemented to find the most convenient reaction mechanism from 15 options (Table 4). After that, the values of the A0 can be obtained from the slope of the linear relationships of Equations (3)–(7) when the reaction mechanism has been specified.

2.4. Thermodynamic Parameter Analysis of PET Pyrolysis

The various thermodynamic model parameters for the PET pyrolysis reaction based on the model fittings’ calculated values of (Ea, A0, and Tp), can be determined by the following three equations:
H = E a R T p
G = E a + R T p l n k B T p h A
S = H G T p
where: ∆H is the change in enthalpy, ∆G is the change in Gibbs free energy, ∆S is the change in entropy, Tp is the maximum peak temperature obtained from the derivative thermogravimetric curves, kB is the Boltzmann constant (1.381 × 10−23 J/K), and h is the Planck constant (6.626 × 10−34 J/s).
The thermodynamic parameters ( H ,   G ,   a n d   S ) are of great importance to the optimization of the large-scale reactor used for pyrolysis. In addition, it is important to verify the energy and the suitability of PET pyrolysis process using the thermodynamic parameters (Dhyani et al. (2017) [22]).

3. Results and Discussion

3.1. TGA of PET

The thermogravimetric (TG) and derivative thermogravimetric (DTG) curves of the PET pyrolysis at 2, 5, 10, and 20 K min−1 heating rates are shown in Figure 1a,b. All curves show similarities in their trend, with shifting to the right (higher temperatures) as the heating rate increases and the mass loss at constant temperature decreases (Figure 1a). A higher heating rate means more energy will be added to the sample, pushing the process to occur at a higher rate and temperature. In addition, an increased heating rate may alter the kinetics of the PET pyrolysis process, which can change the characteristic temperatures (Chowdhury et al. (2023) [16]). Additionally, with the increasing heating rate, the DTG peak and mass loss rate also increased (Figure 1b). This behavior may be due to the heat transfer limitation or thermal lag (Al-Salem et al. (2017) [23]).
As shown in Figure 1, the pyrolysis of PET occurs in the temperature ranges of 650 to 750 K with about 20 wt% residue production. These two curves showed only one reaction stage, as reported elsewhere (Table 5). The characteristic temperatures of pyrolysis are presented in Table 5. This indicates that the characteristic temperatures (onset, peak, and final) found in this study were in agreement with many previous researchers’ reported work (Table 5). Table 5 also shows production of different residue amounts during PET pyrolysis, as reported by many researchers. PET pyrolysis represents the cross-linking of the products to produce more polyaromatics, responsible for char formation (Singh et al. (2020) [24]). Ganeshan et al. (2018) [12] observed 2 stages of reaction, with the first main initial decomposition occurring within 600–740 K with more than 80% mass loss.

3.2. Determination of Kinetic Parameters by Model-Free Methods

Five model-free models (FR, FWO, KAS, STK, and DAEM) were used here to calculate the value of Ea from the TGA data set and from the slope of the fitting plots as shown in Table 1. Regression of all plots is also shown in Figure 2, while the values of Ea and the obtained R2 value by the five models are displayed in Figure 3 and tabulated in Table 6 at a conversion range of 0.1–0.8. A low correlation (R2) value at a reaction conversion of 0.9 indicates inapplicability of this fitted model (Mishra et al. (2019) [15], Damartzis et al. (2011) [25]). The average activation energies obtained from FR, FWO, KAS, STK, and DARM methods were 204, 220, 211, 215, and 214 kJ mol−1, respectively. The correlation coefficient (R2) was found to be higher than 0.8 for most of the methods except at 2 conversion reactions (0.7 and 0.8) for the FR method, where the values of R2 were between 0.7 and 0.8. This low correlation coefficient value may be due to the end of the reaction.
The obtained Ea values from all curve-fitting methods except the FR method were in very close agreement, indicating the superior applicability of all four other models compared to the FR method (Mishra et al. (2019) [15]). The final average activation value of 212 kJ mol−1 for the five methods was very close to the values obtained by many other researchers (Table 7). For example, Osman et al. (2020) [13] reported Ea values of 165–195 kJ mol−1 by differential isoconversional methods and 166–180 kJ mol−1 by FWO. Chowdhury et al. (2023) [16] noted a very low value of Ea (3.31–8.79 kJ mol−1) using the FR model, albeit with reasonably high values of R2 (0.8648–0.9567). Das and Tiwari (2019) [14] calculated Ea by advance isoconversional method (AIC) at higher heating ranges (5–50 K min−1), and its value was found to be within the range of 203–355 kJ mol−1, with large variation over the conversion ranges of 0.05–0.8. They observed that the change of Ea tended to start at a low value and increase over the duration of the pyrolysis reaction. Mishra et al. (2019) [15] obtained variation values of Ea (KAS = 210–241 kJ mol−1, FWO = 211–241 kJ mol−1, STK = 211–242 kJ mol−1, and FR = 208.6–236.0 kJ mol−1) over reaction conversion of 0.1–0.8.

3.3. Determining the Kinetic Parameters by Model-Fitting Methods

The CR model, Equation (8), was applied to identify the suitable reaction mechanism/s for PET pyrolysis. Therefore, values of Ea and A0 at different heating rates for 15 solid-state reaction mechanisms were determined from the linear fitting plots between ln(g(α)/T2) versus 1/T; the obtained kinetic parameters are presented in Table 8. As shown in Table 8, the CR method was well fitted for the TGA data of the PET pyrolysis, with a higher linear regression coefficient of R2 > 0.99.
Only one reaction demonstrated a straight line, and the values of Ea calculated by CR with a function of g(α) (F1–P4) are shown in Table 8. A large deviation in the value of Ea was found (in the range of 45–503 kJ min−1) in the case of different reaction mechanisms (F1–P4) that did not suit the PET pyrolysis reaction.
The correct value for these ranges of activations will be selected from the comparison between the right and left side of the Criado (shown in Figure 4 and Table 9) method. Most of the values of Ea, except F1, were not considered because of their large deviation from the expected Ea values. All these graphs, except F1, were deleted from Figure 4a,c,e,g because they were not close agreement with the experimental range values. Figure 4b,d,f,h show only the most controlling model reaction mechanism (reaction order models—first order—F1) (Das and Tiwari (2019) [13]). Table 9 lists Ea, ln (A0), and R2 for all four tests. The “g(α)-F1”, as the final reaction mechanism, will be used with each of five model-free isoconversional methods (FR, FWO, KAS, STK, and DAEM) to obtain the value of ln(A0) (Table 10).
Some studies in the literature reported the direct use of the CR method alone, based on using one mechanism step prior to the model reaction. However, in this current research, the appropriate mechanistic model as a second trend was determined by the Criado model’s fitting approach. The CR model’s approach to a single heating rate could cause a high possibility of data failure (Chowdhury et al. (2023) [16]). Chowdhury et al. (2023) [16] used two nonisoconversional models (CR and Arrhenius) and found an increase in Ea values with an increasing heating rate; this enhancement may be attributed to the increase in the reaction rate. They found abnormal values of Ea (CR = 1.02 × 104–1.05 × 104 kJ mol−1, Arrhenius = 888.75–1889.94 kJ mol−1). Ganeshan et al. (2018) [12] applied the CR method alone to obtain Ea and A0, assuming reaction orders between 0.1 and 3 and selecting the best R2 values at each heating rate. They observed that Ea decreased with the increased in heating rate, which was not observed in our study. They found that the best value of the reaction order was 1.5 with Ea = 244.496 kJ min−1 and R2 = 0.997 for PET pyrolysis.
Das and Tiwari (2019) [14] also determined the reaction mechanism by selecting the best linear fitting curves for each heating rate between the theoretical and the experimental value. They detected that all the models with equal linearity coefficient R2 were represented as A2, A3, A4, and F1. Then, among the four models, F1 was identified as the best model, producing the lowest root-mean-square error (RMSE).
The selected mechanism by the Criado method or compensation effect (Equation (9)) could be used to determine the linearity between lnA0 and Ea. Figure 5 shows the linear relationship (R2 = 0.9991). This improves the convenience of the suggested model for PET pyrolysis. Das and Tiwari (2019) [14] confirmed that the values A0 can compensate for the values of Ea by obtaining a straight line from a plot of l n A 0 with Ea over the entire conversion range.

3.4. Thermodynamic Parameter Analysis

The thermodynamic parameters ( H ,   G ,   a n d   S ) were calculated for (for 2, 5, 10, and 20 K min−1) and are shown Table 11.
Table 11 shows positive values of H (206.34, 206.3, 206.1, and 206.0 kJ mol−1 at 2, 5, 10, and 20 K min−1, respectively), indicating that the main reaction was endothermic in nature. A similar observation was reported by Chowdhury et al. (2023) [16].
The positive and negatives values of G and S indicate that the PET pyrolysis was a nonspontaneous process in nature. The negative signs of ∆S at different heating rates indicate that the process became less disturbing in thermodynamic equilibrium and low in reactivity (Chowdhury et al. (2023) [16]). The thermodynamic results reveal the promising potential of PET pyrolysis to be efficiently used to produce bioenergy. Chowdhury et al. (2023) [16] applied the kinetics parameters of the CR model to determine thermodynamic parameter values at different heating rates.
As long as pyrolysis is a process used to produce high-energy fuel, our comprehension of changes in enthalpy, entropy, and free energy with different heating rates must consider thermodynamics (Enyoh et al. (2022) [28], Xu and Chen (2013) [29]). Chowdhury et al. (2023) [16] computed the thermodynamic parameters using the kinetics parameters obtained by the CR and Arrhenius models.

4. Conclusions

Pyrolysis of PET at various heating rates of 2, 5, 10, and 20 K min−1 was investigated and the results are discussed here. TG and DTG curves indicated that PET pyrolysis occurred in 1 stage covering the temperature range of 650–750 K. TGA experimental data were fitted well by the five most-popular (differential and integral) model-free isoconversional models (FR, FWO, KAS, STK, and DAEM), and the values of Ea and the pre-exponential factor for PET pyrolysis were determined. However, the most appropriate reaction mechanism was fixed by the CR nonisoconversional model and the master plot of the Criado method. The appropriateness of the mechanisms was shown by the linearity of the relation between lnA0 and Ea.
The thermodynamic properties for PET pyrolysis reaction at different heating rates showed that this reaction was endothermic in nature and confirmed the suitability for the production of bioenergy by the pyrolysis process.

Author Contributions

Conceptualization, I.D.; data curation Z.A. and I.D.; formal analysis, I.D.; funding acquisition, Z.A. and I.D.; investigation, I.D.; methodology, I.D.; project administration, software, Z.A. and I.D.; validation, Z.A. and I.D.; visualization, I.D.; writing—original draft, I.D.; writing—review and editing, Z.A. and I.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC were funded by the Deputyship for Research and Innovation, Ministry of Education of Saudi Arabia, through project number INST059.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Unavailable due to privacy or ethical restrictions.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education, Saudi Arabia, for funding this research work through project number INST059.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martín-Gullón, I.; Esperanza, M.; Font, R. Kinetic model for the pyrolysis and combustion of poly-(ethylene terephthalate) (PET). J. Anal. Appl. Pyrolysis 2001, 58, 635–650. [Google Scholar] [CrossRef]
  2. Diaz-Silvarrey, L.S.; McMahon, A.; Phan, A.N. Benzoic acid recovery via waste poly(ethylene terephthalate) (PET) catalytic pyrolysis using sulphated zirconia catalyst. J. Anal. Appl. Pyrolysis 2018, 134, 621–631. [Google Scholar] [CrossRef] [Green Version]
  3. Moltó, J.; Font, R.; Conesa, J.A. Kinetic model of the decomposition of a PET fibre cloth in an inert and air environment. J. Anal. Appl. Pyrolysis 2007, 79, 289–296. [Google Scholar] [CrossRef]
  4. Çepelioğullar, Ö.; Pütün, A. Thermal and kinetic behaviors of biomass and plastic wastes in co-pyrolysis. Energy Convers. Manag. 2013, 75, 263–270. [Google Scholar] [CrossRef]
  5. Yang, J.; Miranda, R.; Roy, C. Using the DTG curve fitting method to determine the apparent kinetic parameters of thermal decomposition of polymers. Polym. Degrad. Stab. 2001, 73, 455–461. [Google Scholar] [CrossRef]
  6. Saha, B.; Maiti, A.K.; Ghoshal, A.K. Model-free method for isothermal and non-isothermal decomposition kinetics analysis of PET sample. Thermochim. Acta 2006, 444, 46–52. [Google Scholar] [CrossRef]
  7. Brems, A.; Baeyens, J.; Vandecasteele, C.; Dewil, R. Polymeric Cracking of Waste Polyethylene Terephthalate to Chemicals and Energy. Technol. Pap. 2011, 61, 721–731. [Google Scholar] [CrossRef]
  8. Diaz Silvarrey, L.S.; Phan, A.N. Kinetic study of municipal plastic waste. Int. J. Hydrogen Energy 2016, 41, 16352–16364. [Google Scholar] [CrossRef] [Green Version]
  9. Miandad, R.; Barakat, M.A.; Rehan, M.; Aburiazaiza, A.S.; Ismail, I.M.I.; Nizami, A.S. Plastic waste to liquid oil through catalytic pyrolysis using natural and synthetic zeolite catalysts. Waste Manag. 2017, 69, 66–78. [Google Scholar] [CrossRef]
  10. Dimitrov, N.; Krehula, L.K.; Siročić, A.P.; Hrnjak-Murgić, Z. Analysis of recycled PET bottles products by pyrolysis-gas chromatography. Polym. Degrad. Stab. 2013, 98, 972–979. [Google Scholar] [CrossRef]
  11. Chandrasekaran, S.R.; Kunwar, B.; Moser, B.R.; Rajagopalan, N.; Sharma, B.K. Catalytic thermal cracking of postconsumer waste plastics to fuels. 1. Kinetics and optimization. Energy Fuels 2015, 29, 6068–6077. [Google Scholar] [CrossRef]
  12. Ganeshan, G.; Shadangi, K.P.; Mohanty, K. Degradation Kinetic Study of Pyrolysis and Co-Pyrolysis of Biomass with Polyethylene Terephthalate (PET) Using Coats–Redfern Method. J. Therm. Anal. Calorim. 2018, 131, 1803–1816. [Google Scholar] [CrossRef]
  13. Osman, A.I.; Farrell, C.; Al-Muhtaseb, A.H.; Al-Fatesh, A.S.; Harrison, J.; Rooney, D.W. Pyrolysis kinetic modelling of abundant plastic waste (PET) and in-situ emission monitoring. Environ. Sci. Eur. 2020, 32, 112. [Google Scholar] [CrossRef]
  14. Das, P.; Tiwari, P. Thermal degradation study of waste polyethylene terephthalate (PET) under inert and oxidative environments. Thermochim. Acta 2019, 679, 178340. [Google Scholar] [CrossRef]
  15. Mishra, R.K.; Sahoo1, A.; Mohanty, K. Pyrolysis kinetics and synergistic effect in co-pyrolysis of Samanea saman seeds and polyethylene terephthalate using thermogravimetric analyser. Bioresour. Technol. 2019, 289, 121608. [Google Scholar] [CrossRef]
  16. Chowdhury, T.; Wang, Q. Study on Thermal Degradation Processes of Polyethylene Terephthalate Microplastics Using the Kinetics and Artificial Neural Networks Models. Processes 2023, 11, 496. [Google Scholar] [CrossRef]
  17. Dubdub, I.; Al-Yaari, M. Pyrolysis of mixed plastic waste: I. Kinetic study. Materials 2020, 13, 4912. [Google Scholar] [CrossRef] [PubMed]
  18. Dubdub, I.; Alhulaybi, Z. Catalytic Pyrolysis of PET Polymer Using Nonisothermal Thermogravimetric Analysis Data: Kinetics and Artificial Neural Networks Studies. Polymers 2023, 15, 70. [Google Scholar] [CrossRef]
  19. Koga, N.; Vyazovkin, S.; Burnham, A.K.; Favergeon, L.; Muravyev, N.V.; Perez-Maqueda, L.A.; Saggese, C.; Sánchez-Jiménez, P.E. ICTAC Kinetics Committee Recommendations for Analysis of Thermal Decomposition Kinetics. Thermochim. Acta 2023, 719, 179384. [Google Scholar] [CrossRef]
  20. Aboulkas, A.; El Harfi, K.; El Bouadili, A. Thermal degradation behaviors of polyethylene and polypropylene. Part I Pyrolysis kinetics and mechanisms. Energy Convers. Manag. 2010, 51, 1363–1369. [Google Scholar] [CrossRef]
  21. Dubdub, I. Kinetics Study of Polypropylene Pyrolysis by Non-Isothermal Thermogravimetric Analysis. Materials 2023, 16, 584. [Google Scholar] [CrossRef] [PubMed]
  22. Dhyani, V.; Kumar, J.; Bhaskar, T. Thermal Decomposition Kinetics of Sorghum Straw via Thermogravimetric Analysis. Bioresour. Technol. 2017, 245, 1122–1129. [Google Scholar] [CrossRef]
  23. Al-Salem, S.M.; Antelava, A.; Constantinou, A.; Manos, G.; Dutta, A. A review on thermal and catalytic pyrolysis of plastic solid waste (PSW). J. Environ. Manag. 2017, 197, 177–198. [Google Scholar] [CrossRef] [PubMed]
  24. Singh, R.K.; Ruj, B.; Sadhukhan, A.K.; Gupta, A. A TG-FTIR investigation on the co-pyrolysis of the waste HDPE, PP, PS and PET under high heating conditions. J. Energy Inst. 2020, 93, 1020–1035. [Google Scholar] [CrossRef]
  25. Damartzis, T.; Vamvuka, D.; Sfakiotakis, S.; Zabaniotou, A. Thermal degradation studies and kinetic modeling of cardoon (Cynara cardunculus) pyrolysis using thermogravimetric analysis (TGA). Bioresour. Technol. 2011, 102, 6230–6238. [Google Scholar] [CrossRef]
  26. Senneca, O.; Chirone, R.; Salatino, P. Oxidative pyrolysis of solid fuels. J. Anal. Appl. Pyrolysis 2004, 71, 959–970. [Google Scholar] [CrossRef]
  27. Girija, B.G.; Sailaja, R.R.N.; Madras, G. Thermal degradation and mechanical properties of PET blends. Polym. Degrad. Stab. 2005, 90, 147–153. [Google Scholar] [CrossRef]
  28. Enyoh, C.E.; Wang, Q. Combined Experimental and Molecular Dynamics Removal Processes of Contaminant Phenol from Simulated Wastewater by Polyethylene Terephthalate Microplastics. Environ. Technol. 2022, 1–20. [Google Scholar] [CrossRef]
  29. Xu, Y.; Chen, B. Investigation of Thermodynamic Parameters in the Pyrolysis Conversion of Biomass and Manure to Biochars Using Thermogravimetric Analysis. Bioresour. Technol. 2013, 146, 485–493. [Google Scholar] [CrossRef]
Figure 1. TG (a) and DTG (b) curves of PET pyrolysis at different heating rates.
Figure 1. TG (a) and DTG (b) curves of PET pyrolysis at different heating rates.
Polymers 15 03010 g001
Figure 2. Regression lines of the experimental data of PET pyrolysis by (a) FR, (b) FWO, (c) KAS (d) STK, and (e) DAEM models.
Figure 2. Regression lines of the experimental data of PET pyrolysis by (a) FR, (b) FWO, (c) KAS (d) STK, and (e) DAEM models.
Polymers 15 03010 g002
Figure 3. Regression lines of the experimental data of PET pyrolysis.
Figure 3. Regression lines of the experimental data of PET pyrolysis.
Polymers 15 03010 g003
Figure 4. Master plots of different kinetic models and experimental data of four tests; (a,b)PET2, (c,d) PET5, (e,f) PET10, and (g,h) PET20.
Figure 4. Master plots of different kinetic models and experimental data of four tests; (a,b)PET2, (c,d) PET5, (e,f) PET10, and (g,h) PET20.
Polymers 15 03010 g004aPolymers 15 03010 g004b
Figure 5. Linear fitted curve for the compensation effect.
Figure 5. Linear fitted curve for the compensation effect.
Polymers 15 03010 g005
Table 1. The proximate and ultimate analysis of PET (Dubdub and Alhulaybi (2023) [18]).
Table 1. The proximate and ultimate analysis of PET (Dubdub and Alhulaybi (2023) [18]).
Proximate Analysis, wt%Ultimate Analysis, wt%
MoistureVolatileAshCHNO
0.52388.23111.24664.2564.367031.377
Table 2. Equations for model-free methods (Dubdub (2023) [20]).
Table 2. Equations for model-free methods (Dubdub (2023) [20]).
MethodEquationIntegral (I) or
Differential (D)
Plot
FR ln β d α d T = ln [ A 0 f ( α ) ] E a R T (3)D ln β d α d T   v s .   1 T
FWO ln β = l n A 0 E a R g ( α ) 5.331 1.052 E a R T (4)I ln β   v s .   1 T
KAS ln β T 2 = l n A 0 R E   g ( α ) E a R T (5)I ln β / T 2   v s .     1 T
STK l n β T 1.92 =   l n A 0 E a R g ( α ) 1.0008 E a R T (6)I ln β / T 1.92   v s .     1 T
DAEM l n β T 2 =   l n A 0 R E a + 0.6075 E a R T (7)I ln β / T 2   v s .     1 T
Table 3. Equations for model-fitting methods (Dubdub (2023) [21]).
Table 3. Equations for model-fitting methods (Dubdub (2023) [21]).
MethodEquation
CR l n g ( α ) T 2 = l n A 0 R β E E R T (8)
Criado Z ( α ) Z ( 0.5 ) = f α g ( α ) f 0.5 g ( 0.5 ) = T α T 0.5 2 ( d α d t ) α ( d α d t ) 0.5 (9)
Table 4. Solid-state thermal reaction mechanism (Dubdub (2023) [21]).
Table 4. Solid-state thermal reaction mechanism (Dubdub (2023) [21]).
Reaction MechanismCode f ( α ) g ( α )
Reaction order models—first orderF11 − α l n ( 1 α )
Reaction order models—second orderF2 ( 1 α ) 2 ( 1 α ) 1 1
Reaction order models—third orderF3 ( 1 α ) 3 [ ( 1 α ) 1 1 ] / 2
Diffusion model—one dimension D1 1 / 2 α 1 α 2
Diffusion model—two dimensions D2 [ l n ( 1 α ) ] 1 1 α ln 1 α + α
Diffusion model—three dimensions D3 3 / 2 [ 1 ( 1 α ) 1 / 3 ] 1 [ 1 ( 1 α ) 1 / 3 ] 2
Nucleation models—two dimensionsA2 2 ( 1 α ) [ ln 1 α ] 1 / 2 [ ln 1 α ] 1 / 2
Nucleation models—three dimensionsA3 3 ( 1 α ) [ ln 1 α ] 1 / 3 [ ln 1 α ] 1 / 3
Nucleation models—four dimensions A4 4 ( 1 α ) [ ln 1 α ] 1 / 4 [ ln 1 α ] 1 / 4
Geometrical contraction models—one dimension R11 α
Geometrical contraction models—sphere R2 2 ( 1 α ) 1 / 2 1 - ( 1 α ) 1 / 2
Geometrical contraction models—cylinderR3 3 ( 1 α ) 1 / 3 1 - ( 1 α ) 1 / 3
Nucleation models—two-power lawP2 2 α 1 / 2 α 1 / 2
Nucleation models—three-power lawP3 3 α 2 / 3 α 1 / 3
Nucleation models—four-power lawP4 4 α 3 / 4 α 1 / 4
Table 5. Comparison of pyrolysis characteristics of PET at different heating rates between current work and earlier reported studies by various investigators.
Table 5. Comparison of pyrolysis characteristics of PET at different heating rates between current work and earlier reported studies by various investigators.
Heating Rate K/minThis WorkPrevious WorkReferences
Onset
(K)
Peak
(K)
Final
(K)
Mass
Loss (%)
Onset
(K)
Peak
(K)
Final
(K)
Mass Loss
(%)
262068069080623667694100Osman et al. (2020) [13]
565068571085658
623
700
673
723
733
80
80
Das and Tiwari (2019) [14]
Diaz Silvarrey and Phan (2016) [8]
1067071072580671
643
633
643
585
711
714
700
693
648
707
748
775
773
743
856
80
80
79.78
80
27.5
Das and Tiwari (2019) [14]
Yang et al. (2001) [5]
Çepelioğullar and Pütün (2013) [4]
Diaz Silvarrey and Phan (2016) [8]
Mishra et al. (2019) [15]
Chowdhury et al. (2023) [16]
2068072575095681
673
603
721
703
713
759
773
882
80
80
22.5
Das and Tiwari (2019) [14]
Diaz Silvarrey and Phan (2016) [8]
Chowdhury et al. (2023) [16]
30 661672
716
87827.5Mishra et al. (2019) [15]
Chowdhury et al. (2023) [16]
40 690
698
693
733
748
723
783
798
803
80
82
80
Das and Tiwari (2019) [14]
Singh et al. (2020) [24]
Diaz Silvarrey and Phan (2016) [8]
50 698743
684
79380Das and Tiwari (2019) [14]
Mishra et al. (2019) [15]
Table 6. Activation energy values obtained by five model-free methods.
Table 6. Activation energy values obtained by five model-free methods.
ConversionFRFWOKASSTKDAEMAverage
E
kJ mol−1
R2E
kJ mol−1
R2E
kJ mol−1
R2E
kJ mol−1
R2E
kJ mol−1
R2E
kJ mol−1
R2
0.12130.87772360.90052250.89142260.89182250.89142250.8899
0.22140.88582290.8982180.88822180.88862180.88822190.8912
0.32090.90332270.90472150.89522160.89572150.89522160.8986
0.42000.84212240.90372130.8942130.89442130.8942130.8838
0.52000.83422210.89532090.88462100.8952090.88462100.8727
0.62040.83142190.88482070.85462360.85512360.85462200.8571
0.71990.75832170.87252050.85942060.85992050.85942060.8340
0.81900.76461860.83661970.81981970.82061970.81981930.8104
Average2040.83722200.88702110.87342150.87512140.87342120.8692
Table 7. Activation energies from different published papers.
Table 7. Activation energies from different published papers.
ReferencesE (kJ mol−1)Method
Yang et al. (2001) [5]242Curve fitting
Senneca et al. (2004) [26]217FR
Girij et al. (2005) [27]227
208
236
FR
Ozawa
Kissinger
Saha et al. (2006) [6]180–208VY
Das and Tiwari (2019) [14]203–355AIC
Osman et al. (2020) [13]165.6
166–180
165–195
ASTME698
FWO
FR
Mishra et al. (2019) [15]225.64
230.71
230.55
231.03
FR
KAS
FWO
STK
Table 8. Kinetic parameters obtained by CR model.
Table 8. Kinetic parameters obtained by CR model.
Reaction MechanismCodePET2PET5
Ea
kJ mol−1
ln(A0)R2Ea
kJ mol−1
ln(A0)R2
Reaction order models—first orderF124641.30.999627747.510.9999
Reaction order models—second orderF228648.810.998939368.730.9988
Reaction order models—third orderF333057.010.99853394.140.9966
Diffusion models—one dimensionD143273.870.999738465.650.9983
Diffusion models—two dimensionsD245477.430.999743774.580.9992
Diffusion models—three dimensions D347980.490.999750084.550.9998
Diffusion models—four dimensionsD446577.450.999745873.660.9995
Nucleation models—two dimensions A211817.910.999513321.510.9999
Nucleation models—three dimensions A3759.820.99958512.640.9999
Nucleation models—four dimensions A45313.120.99946112.930.9999
Geometrical contraction models—one-dimension phase boundaryR121034.490.999718630.750.9982
Geometrical contraction models—sphere R222839.110.999722937.890.9995
Geometrical contraction models—cylinderR323437.860.999724440.340.9997
Nucleation models—two-power lawP210014.410.99968812.910.998
Nucleation models—three-power lawP36311.740.99965514.060.9977
Nucleation models—four-power lawP44414.60.99953816.540.9973
Reaction MechanismCodePET10PET20
Ea
kJ mol−1
ln(A0)R2Ea
kJ mol−1
ln(A0)R2
Reaction order models—first orderF126445.050.999621335.330.9999
Reaction order models—second orderF237364.50.997622838.170.9996
Reaction order models—third orderF350387.720.994924441.130.999
Diffusion models—one dimensionD137162.68999640867.961
Diffusion models—two dimensionsD242070.90.9999417691
Diffusion models—three dimensions D347979.970.999942769.31
Diffusion models—four dimensionsD444072.9142168.11
Nucleation models—two dimensions A212720.580.999610115.960.9999
Nucleation models—three dimensions A38112.220.99966314.570.9999
Nucleation models—four dimensions A45814.370.99954517.340.9999
Geometrical contraction models—one-dimension phase boundaryR118029.580.999619832.621
Geometrical contraction models—sphere R221936.12120533.271
Geometrical contraction models—cylinderR323438.350.999920833.321
Nucleation models—two-power lawP28412.630.99959314.561
Nucleation models—three-power lawP35215.340.99945815.371
Nucleation models—four-power lawP43617.670.99934117.911
Table 9. Activation energy of (CR and Criado).
Table 9. Activation energy of (CR and Criado).
Test No.Ea
kJ mol−1
ln(A0)R2Reaction Mechanism
PET224641.30.9996Reaction order models—first order—F1
PET527747.510.9999Reaction order models—first order—F1
PET1026445.050.9996Reaction order models—first order—F1
PET2021335.330.9999Reaction order models—first order—F1
Table 10. Pre-exponential factor values obtained by isoconversional models.
Table 10. Pre-exponential factor values obtained by isoconversional models.
Conversion ln[A0 (min−1)]
FRFWOKASSTKDAEMAverage
0.135.4537.5617.2717.8739.3329.50
0.235.8436.3115.9816.5837.2228.39
0.334.9435.9915.6516.2536.427.85
0.433.4635.5915.2315.8435.627.14
0.533.6335.0514.6915.2834.7126.67
0.634.3934.7419.4419.9139.329.56
0.733.5334.4914.0914.6933.5326.07
0.839.5133.1812.7713.3831.8326.13
Average35.0935.3615.6416.2335.9927.66
Table 11. Thermodynamic parameters.
Table 11. Thermodynamic parameters.
Heating Rates (K/min)251020
Kinetic Parameters
Ea (kJ/mol)212
A (min−1)1.29 × 1011
Tp (K)680685710725
Thermodynamic Parameters
H (kJ/mol)206.34206.3206.1206.00
G (kJ/mol)215.4215.48215.82216.03
S (kJ/mol.K)−0.01332−0.0134−0.01369−0.01383
Potential Energy Barrier
Ea–∆H (kJ/mol) *5.82
* Based on the mean values of ∆H.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alhulaybi, Z.; Dubdub, I. Comprehensive Kinetic Study of PET Pyrolysis Using TGA. Polymers 2023, 15, 3010. https://doi.org/10.3390/polym15143010

AMA Style

Alhulaybi Z, Dubdub I. Comprehensive Kinetic Study of PET Pyrolysis Using TGA. Polymers. 2023; 15(14):3010. https://doi.org/10.3390/polym15143010

Chicago/Turabian Style

Alhulaybi, Zaid, and Ibrahim Dubdub. 2023. "Comprehensive Kinetic Study of PET Pyrolysis Using TGA" Polymers 15, no. 14: 3010. https://doi.org/10.3390/polym15143010

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