Next Article in Journal
Benzylidene Cyclopentanone Derivative Photoinitiator for Two-Photon Photopolymerization-Photochemistry and 3D Structures Fabrication for X-ray Application
Previous Article in Journal
Effects of Accelerating the Ageing of 1D PLA Filaments after Fused Filament Fabrication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Pyrolysis of PET Polymer Using Nonisothermal Thermogravimetric Analysis Data: Kinetics and Artificial Neural Networks Studies

Department of Chemical Engineering, King Faisal University, Al-Ahsa 31982, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(1), 70; https://doi.org/10.3390/polym15010070
Submission received: 7 December 2022 / Revised: 16 December 2022 / Accepted: 21 December 2022 / Published: 24 December 2022
(This article belongs to the Section Polymer Physics and Theory)

Abstract

:
This paper presents the catalytic pyrolysis of a constant-composition mixture of zeolite β and polyethylene terephthalate (PET) polymer by thermogravimetric analysis (TGA) at different heating rates (2, 5, 10, and 20 K/min). The thermograms showed only one main reaction and shifted to higher temperatures with increasing heating rate. In addition, at constant heating rate, they moved to lower temperatures of pure PET pyrolysis when a catalyst was added. Four isoconversional models, namely, Kissinger–Akahira–Sunose (KAS), Friedman, Flynn–Wall–Qzawa (FWO), and Starink, were applied to obtain the activation energy (Ea). Values of Ea acquired by these models were very close to each other with average value of Ea = 154.0 kJ/mol, which was much lower than that for pure PET pyrolysis. The Coats–Redfern and Criado methods were employed to set the most convenient solid-state reaction mechanism. These methods revealed that the experimental data matched those obtained by different mechanisms depending on the heating rate. Values of Ea obtained by these two models were within the average values of 157 kJ/mol. An artificial neural network (ANN) was utilized to predict the remaining weight fraction using two input variables (temperature and heating rate). The results proved that ANN could predict the experimental value very efficiently (R2 > 0.999) even with new data.

Graphical Abstract

1. Introduction

Plastic waste has become a serious global problem, especially in countries where a huge quantity of plastic waste is produced and disposed. There are many academic research centers attempting to find a suitable way to recycle such vast waste [1].
PET is one of six polymers, with the others being low-density polyethylene (LDPE), high-density polyethylene (HDPE), polystyrene (PS), polypropylene (PP), and polyvinyl chloride (PVC). PET accounts for almost 7.6 wt% of plastic solid waste (PSW) [2,3]. It is very commonly used in manufacturing food packaging and fiber cloth and is considered as the most recycled material [4]. PET is mostly used in manufacturing fiber, staple filaments, textile, and carpet [3,5].
TGA is a thermal analysis method that is best known for calculating the kinetic parameters of plastic pyrolysis. Estimation of kinetic parameters from TGA data can be performed by different methods. These methods use either multiple TGA datasets at different heating rates (called “model-free” or “isoconversional” method) or one single TGA dataset (called “model-fitting” method) and generate the final kinetic equation using either differential method or integral technique.
In the following literature survey, only the kinetics of PET by TGA data and the kinetics of catalytic cracking of different plastic waste polymers by TGA data are examined. Yang et al. (2001) [6] initiated the kinetic study of PET using the Arrhenius equation with Friedman’s method for DTG curves to obtain the kinetic parameters of PET polymers. TG and DTG curves showed only one single decomposition that started at 370 °C, reached maximum at 441 °C, and ended at 502 °C at 10 °C/min heating rate. The authors obtained activation energy value of 242 kJ/mol with reaction order (n) of 1. Martín-Gullón et al. (2001) [2] showed that the pyrolysis reaction could almost be divided into two reactions. The first one with 80% of weight loss, considered as the main reaction within the range of the temperature (700–750 K), had activation energy of 256.4 kJ/mol, while the second one with only 6% weight loss had activation energy of 332.3 kJ/mol. Girija et al. (2005) [7] highlighted that PET had only single-stage degradation at temperature of 440 °C with weight loss of 60%. Saha and Ghoshal (2005) [8] studied the pyrolysis kinetics for different soft drink bottles to obtain the activation energy value using different models. They reported Ea within the values of 320–339.89 kJ/mol.
Aguado et al. (2007) [9] measured the activity of two catalysts, namely, n-HZMS-5 and Al-MCM-41, on the degradation of LDPE. This activity was measured by how much the TG shifted to lower temperature. They stated that the pure polymer started to degrade at 477 °C and shifted to 419 °C with Al-MCM-41 and then further to 396 °C with zeolite n-HZSM-5. They attributed this strong activity for zeolite n-HZSM-5, which has weak acid properties, to the large pore dimension. Moltó et al. (2007) [10] investigated the kinetics of PET fiber cloth pyrolysis with three different heating rates. They found three different partial weight loss reactions, with the second reaction being the main one within the range of 640–730 K temperature. They first suggested three independent reactions with activation energy values of 121.5, 238.3, and 300.0 kJ/mol but later considered the second reaction with 238.3 kJ/mol as the main reaction. These results were close to the findings of Yang et al. (2001) [6], who showed an energy value of 242 kJ/mol.
Halász et al. (2009) [11] repeated the same experiment carried out by Aguado et al. (2007) [9] but with PET polymer. They found that pure polymer PET had DTG peak maximum equal to 473 °C, which shifted to 412 °C with H-ZSM-5 and to 458 °C with Ti-MCM-41.
Çepelioğullar and Pütün (2013) [12] investigated the effect of different biomasses (cotton stalk (CS), hazelnut shell (HS), sunflower residue (SFR), and arid land plant Euphorbia rigida (ER)) in the copyrolysis with PET. They mentioned that the pyrolysis of PET decomposed at two stages, which started at 360 °C with peak temperature of 427.7 °C and weight loss of about 79.78%. They made unusual observation when the pyrolysis of biomass–PET mixtures occurred at temperatures higher than pure biomass and PET. Moreover, they found the range of activation energy values of PET was higher than the biomass used. This demonstrated that the decomposition of biomass started at lower temperatures. They attributed this difference to the structural differences between them. For pure PET, they obtained a reaction in two stages. The first stage had temperature range of 373–443 °C, conversion range of 0.4–81%, and activation energy value of 347.4 kJ/mol, while the second stage had a temperature range of 448–503 °C, conversion range of 88–99%, and activation energy value of 172.6 kJ/mol. They presented two tables of activation energy values for different pure biomasses (CS, HS, SFR, and ER), PET, and the biomass–PET mixture. The average activation energy value of CS, PET, and CS–PET decompositions were 71, 260, and 95.7 kJ/mol, respectively.
Lee et al. (2018) [13] studied the catalytic decomposition of mixed torrefied cork oak (TOak) and HDPE using three different catalysts (mesoporous HY1, mesoporous HY2, and microporous HY). They stated that the catalytic pyrolysis of HDPE over HY had two reactions within the range of 200–300 and 300–400 °C with peak temperature of 392 °C, which was lower than the pure FDPE temperature of 477 °C. This behavior was different from the catalytic copyrolysis of Oak/HDPE and TOak/HDPE over HY. In these cases, the peak temperature was shifted higher to 411 and 414 °C, respectively. Diaz Silvarrey and Phan (2016) [14] found that the decomposition of PET by TGA started earlier than most other polymers with a thermal decomposition range of 395–520 °C. They used the KAS method to obtain 197.61 kJ/mol as the average value of activation energy and 4.84 × 1014 s−1 as the pre-exponential factor.
Xiang et al. (2018) [15] investigated the catalytic decomposition of cobalt-modified ZSM-5 catalyst on copyrolysis of lignocellulosic biomass (rice straw (RS)) and linear low-density polyethylene (LLDPE). They stated that there was only one peak for the pyrolysis of pure RS at 341.39 °C and pure LLDPE at 491.53 °C, two peaks for the copyrolysis of RS–LLDPE at 339.57 and 470.06 °C, and two peaks for the pyrolysis of RS–LLDPE with Co/ZSM-5 catalyst at 336.32 and 451.73 °C. They calculated the activation energy of the pyrolysis of LLDPE, RS, RS–LLDPE, and catalytic RS–LLDPE as 212.23, 79.61, 34.27, and 41.04 kJ/mol, respectively. Das and Tiwari (2019) [5] used PET collected from soft drink bottles to evaluate the degradation kinetics of nonisothermal TGA runs with five different heating rates. They obtained activation energy values in the range of 203–355 kJ/mol using the Vyazovkin method and established the F1 reaction model using Criados’ master plot.
Osman et al. (2020) [16] employed TGA analysis for PET pyrolysis with five heating rates with a ratio of 16 between the lowest (0.5 K/min) and highest (8 K/min) rates. They applied different isoconversional methods (integral Friedman, ASTM-E698, and integral FWO) and concluded that the activation energy was between 165 and 195 kJ/mol.
Ali et al. (2020) [17] studied the catalytic cracking kinetics of polystyrene with synthesized copper oxide as a catalyst. They used Friedman, KAS, FWO, and Coats–Redfern to obtain the kinetic parameter values. They concluded that the low value of the activation energy would help to run experiments at lower temperatures. Lai et al. (2021) [18] examined the performance of coal ash (fly and bottom ash) with different blending ratios for the catalytic pyrolysis of LLDPE. They showed that 15 wt% bottom ash had higher effect on catalytic activity compared to 15 wt% fly ash. Yap et al. (2022) [19] used two different catalysts (HZSM-5 zeolite and natural limestone (LS)) for copyrolysis of HDPE and Chlorella vulgaris mixture. They concluded that the combined HZSM-5/LS catalyst had the best effect in the copyrolysis of the HDPE and Chlorella vulgaris mixture, with the activation energy decreasing from 144.93–225.84 to 75.37–76.90 kJ/mol. Kokuryo et al. (2022) [20] studied the catalytic degradation of LDPE using two types of catalysts (zeolite β and zeolite MTES-β) at 5 K/min heating rate. They concluded that zeolite MTES-β pushed LDPE to decompose at a lower temperature than zeolite β. Dourari et al. (2022) [21] and Tarchoun et al. (2022) [22] developed new composites with different materials and tested them using various analytical techniques. They concluded that there was a big drop in the activation energy of the thermal behavior.
In this study, the activation energy of PET obtained by catalytic thermal decomposition was calculated by four model-free methods and two model-fitting techniques. In order to study the effect of the zeolite β catalyst, these kinetic parameter values were compared with the pure PET pyrolysis values collected from the literature survey. Furthermore, the catalytic pyrolysis behavior of PET was determined for the first time by a strongly promising ANN model.

2. Materials and Methods

2.1. Materials

PET polymer samples were collected from Recycled Plastic (Ipoh, Malaysia). The two main analyses of the polymer (shown in Table 1) were performed using PerkinElmer thermal analyzer and PerkinElmer elemental analyzer (Waltham, MA, USA). Zeolite β (its main properties are shown in Table 2) catalyst in powder form was obtained from Alfa Aesar (Ward Hill, MA, USA).
All the details related to preparing the powders and sample weighing can be found in [23].

2.2. Catalytic Thermal Decomposition of PET

The 1020 series TGA7 thermogravimetric analysis was used to meet the objective of this research. Under the control of the system controller, the changes in weight due to chemical reaction were measured. PET polymer pellets were ground into powders before being fed to TGA. A good reproducibility was achieved in the range of 5–100 mg. Therefore, samples weighing 10 mg were used throughout the study. The samples were heated at 2, 5, 10, and 20 K/min under nitrogen at 20 mL/min within the temperature range of 298–873 K at constant composition of zeolite β to PET (20/80 wt/wt).

2.3. Kinetic Theory

The rate of reactions (r) can be expressed as a product of a temperature (T)-dependent function, K(T), and a composition- or conversion (α)-dependent term, f(α), as follows [24]:
r = d α d t = K ( T ) f ( α )
α = w o w w o w f  
The definitions of α, t, T, wo, wo, and wf can be found in [23]. The temperature-dependent term, which is the reaction rate constant (K), is assumed to obey the Arrhenius relationship:
K ( T ) = A 0 exp ( E R T   )
where E is the activation energy (kJ/mol), A0 is the pre-exponential factor K−1, and R is the universal gas constant (8.314 J/mol.K). If it is assumed that a simple nth-order kinetic relationship holds for the conversion-dependent term, then the following applies:
f ( α ) = ( 1 α ) n  
Substituting Equations (3) and (4) in (1) yields the following:
r = d α d t = A 0 exp ( E R T   ) ( 1 α ) n
All the published studies have used either single or multiple thermograms at different heating rates with either differential or integral methods for calculating kinetic parameters from TGA data [24]. However, in this work, two different sets of methods were examined and compared with the published available data. One set used multiple thermograms, called model-free methods, which included Friedman, FWO, KAS, and Starnik (different heating rates and constant conversion). The second set used single thermograms, called model-fitting methods, which included Coat–Redfern and the Criado equation. Derivations of all these equations with the common solid-state thermal reaction mechanisms g(α) used in the Coats–Redfern method can be found in [23,25,26].

2.4. Topology of ANNs

There is a common problem in modeling any engineering process, especially if the final model is very complicated with high nonlinearity and too many parameters. In these cases, ANN can be considered as another alternative option. Muravyev et al. (2021) [27] reviewed the main concept of ANN for application in pyrolysis, thermal analysis, and kinetic studies. They highlighted that there is no single mathematical technique to resolve all pyrolysis problems.
An ANN architecture is ordered in three consecutive layers of neurons: input, hidden, and output. In each layer, there is a weight matrix, bias vector, and output vector [28]. The variables influencing the structure of the network should be fixed and the whole problem area should be represented outside the ANN. ANN architecture should have a number of layers, with each layer having a transfer function and connections between the layers. The type of problem presented will determine the best architecture. The number of neurons in the hidden layer, the number of hidden layers, the momentum rate, and the learning rate are part of the optimizing parameters.
In this research, the performance of an ANN model was monitored by the following statistical parameters: average correlation factor (R2), root mean square error (RMSE), mean absolute error (MAE), mean bias error (MBE), and correlation coefficient (R) statistical correlations [1,29,30,31].
The remaining mass percentage of PET was measured by the ANN model. There are some advantages and maybe some disadvantages in using ANN. Some of these advantages include the ability to work with linear and nonlinear relationships and learn these relationships directly from the data used, while some of these disadvantages including the need for large memory and computational efforts for fitting [32].

3. Results and Discussion

3.1. TG–DTG Analysis of PET

TG and DTG at different heating rates of the catalytic pyrolysis of PET by zeolite β are shown in Figure 1. The thermograms were similar for all different heating rates. Osman et al. (2022) [33] and Farrell et al. (2021) [34] emphasized that TGA experiments should be run according to the International Confederation for Thermal Analysis and Calorimetry (ICTAC) instructions in order to collect the proper data for calculation of kinetic parameters. Therefore, they used 16 as the ratio between the highest and lowest heating rates. Muravyev and Vyazovkin (2022) [35] reviewed the most proper options that should be chosen as advised by the International Confederation for Thermal Analysis and Calorimetry (ICTAC). ICTAC mainly recommends using model-free methods, where the data should be collected at more than three different heating rates. ICTAC emphasized that TGA experiments should be run according to its instructions in order to collect the proper data for kinetic parameter calculations. Because the TG curve showed only one peak reaction, single-step thermal decomposition as recommended by ICTAC was considered to derive the proper kinetic parameters [36].
Catalytic thermal decomposition onset, peak, and final temperatures, which were obtained from the two curves, moved to higher temperature when the heating rate increased from 2 to 20 K/min. Diaz Silvarrey and Phan (2016) [14] mentioned that as the heating rate increases, the peak of DTG will be bigger.
It has been shown that all plastic polymers except PVC degrade by only one reaction, including PET [5,6]. As shown in Table 3, the onset, peak, and final temperatures of catalytic cracking of PET by zeolite β catalyst were smaller than those of cracking decomposition of pure PET. For example, the onset, peak, and final temperatures of catalytic cracking at 10 K/min were 625, 710, 740 K in this study, while they were 671, 711, 748 K in [5] and 643, 714, 775 K in [6]. Al-Salem et al. (2017) [4] highlighted that catalytic cracking is more promising than thermal degradation as it requires lower energy.

3.2. Model-Free Kinetics Calculation

Equation (5) is considered as the starting equation from which all the other equations can be attained. Four different types of isoconversional methods were implemented to calculate the main kinetic parameter (activation energy (Ea) in kJ/mol). The derivation of these four methods can be found in [23]. Friedman, FWO, KAS and Starnik at conversions ranging from 0.1 to 0.8 and the activation energies values for the four methods are presented in Figure 2 and Table 4. From these, it can be seen that the values of activation energy with the Friedman’s method (average 167.5 kJ/mol) were slightly higher than and totally different from the rest three, where the values were homogeneous with an average of 140.2 kJ/mol. These two values were still much less than what has been published for pure PET pyrolysis. Ea values in the range of 203–355 [5], 165–195 [37], and 185–230 [14] kJ/mol have previously been reported for pyrolysis of PET. Therefore, the lowest activation for PET degradation in the presence of zeolite β catalyst proves the suitability of using a catalyst at lower temperatures. This finding has been supported by many researchers [4,9,11,17,20]. Al-Salem et al. (2017) [4] reviewed and confirmed that a catalyst lowers the required pyrolysis temperature for all polymers. As shown in Table 3, there was no trend in the activation energy, except that the Freidman methods started at lower values (130 kJ/mol at α = 0.1), increased over the period of degradation (188 kJ/mol at α = 0.6), and decreasing until the end (136 kJ/mol at α = 0.8). For FWO, KAS, and Starnik, the trend was the same, with a steady increase from an average of 93.33 kJ/mol (α = 0.1) to an average of 164 kJ/mol (α = 0.7) and then a decrease to 156 kJ/mol (α = 0.8).
Generally, these methods are considered more reliable compared to model-fitting methods as they obtain consistent activation energy values for the nonisothermal data [37,38]. They also help to show the complexity of multireaction as the calculation of the activation energy depends on the conversion [39]. Moreover, Özsin and Pütün (2019) [40] found that the activation energy values for the four different methods depended heavily on the conversion value. From the calculations and graphs, it became clear that there is a need in the future to conduct additional experiments with more heating rates to improve some low values of R2 in Table 4 caused by the ratio of 10 between the highest and lowest heating rates.

3.3. Model-Fitting Kinetics Calculation

Two model-fitting (Criado and Coats–Redfern) methods were implemented to study the kinetics of catalytic cracking of PET. For more details relating to derivation, please refer to [26]. The first model-fitting method (Coats–Redfern) was tested by plotting ln [ g ( α ) T 2 ] against 1/T. The most suitable formula of function g ( α ) common solid-state reaction mechanism was selected and presented in [41,42]. Table S1 lists the kinetic parameter (Ea kJ/mol and lnA) and the average correlation factor (R2). It can be seen from the Table S1 that there were big differences in the Ea between these 16 different reaction mechanisms. In the next stage, these data results were connected to the second model-fitting (Criado method). The Criado method was utilized to calculate the reaction model. In the Criado method, the left-hand side of the Criado equation is called a reduced theoretical curve, and the right-hand side represents the experimental data. To use this method, comparison between these two sides were carried out to help us define the exact kinetic model describing the reaction. Figure S1 shows the curves obtained from the Criado method at different heating rates only for the reaction mechanisms close to the experimental curve. Figure S1 and Table 5 reveal that the experimental data matched first-order reaction (F1) or two dimensional phase interfacial reaction (R2) for 2, 5, and 10 K/min heating rates and (F1) or (R2) or two dimensional nucleation and growth reaction (A2) for 20 K/min. Table 5 reveals that there was a range of activation energy values between 104 to 242 kJ/mol with an average of 157 kJ/mol depending on different reaction mechanisms. Das and Tiwari (2019) [5] showed that PET degradation under nitrogen follows many reaction models (A2, A3, A4, and F1) for all the heating rates.

3.4. Catalytic Pyrolysis Prediction by ANN Model

ANN model was developed to monitor the catalytic pyrolysis of PET. Feedforward with backpropagation was developed to predict the remaining weight fraction based on 983 experimental datasets with heating rate and temperature as the input variables.
The collected data is usually divided into three subsets: training set, validation set, and finally test set [43]. Each set has its own role in the whole process.
The whole data comprising 983 sets were randomly divided into three sets as follows: 70% used for training, 15% used for validation, and 15% used for testing. More details about the data are shown in Table 6. Osman and Aggour (2002) [44] mentioned that collecting large sets of data can lead to a model with high accuracy compared to the original data. To obtain the best performance of ANN structure, it is necessary to fix: (1) the number of hidden layers, (2) transfer function for output and hidden layer/s, and (3) number of neurons in each hidden layer. A comparison was then carried out between different ANN structures with variable number according to the above points.
The value of R was used as the main criterion for judging the most efficient structure to estimate the remaining weight fraction as the output variable for 23 runs. A total of 13 runs out of the 23 runs resulted in high value of R. These 13 runs were moved to the second stage comparison, where 12 new simulation sets were tested again against the value of R. Table 7 shows the 12 new sets that were collected for the four heating rates.
The final and best ANN structure (2-15-15-1), as shown in Figure 3, was selected after the two comparison steps. It had 15 neurons for each two hidden layers (tangsig–tangsig transfer functions) with linear transfer function for the output layer. The efficiency and the accuracy of the ANN output is heavily dependent on the number of neurons in the hidden layers. To avoid underfitting and the overfitting, one must select the number of neurons in such a way that the performance function will eventually obtain the optimum value [43,45]. Al-Wahaibi and Mjalli (2014) [38] concluded that the number of neurons in the hidden layer is a very crucial parameter in predicting the accuracy of the ANN output, and using very few numbers may result in something called underfitting. Qinghua et al. (2018) [40] also pointed out that the number of neurons in the hidden layer may cause training error if the number is too small or overfitting problem if the number is too high. The Levenberg–Marquardt algorithm is commonly suited for this type of network [46]. The ANN model was run with different algorithms, namely, scaled conjugate gradient and Bayesian regularization, to check that the test was relevant. As shown in Figure 4, all the comparative results were close to the diagonal, which proved good agreement between the ANN values (Y-axis) and real values (X-axis).
The performance of the current (2-15-15-1) model in predicting remaining weight fraction was examined by calculating the four statistical parameters. As shown in Table 8, all statistical measurements of deviations were notably low for the best ANN model. This meant the selected ANN model could efficiently predict the output variable within an acceptable limit of error. After training the ANN with 983 sets, the recommended architecture (2-15-15-1) was simulated with 12 new input datasets as delineated in Table 7. In this stage, new input data were entered, and the network produced the new simulated output. Figure 5 compares the simulated ANN with the measured output and Table 9 lists the R2, RMSE, MAE, and MBE values. As can be seen, R2 was very high (>0.9999), while RMSE, MAE, and MBE were very low. Figure 6 presents the error histogram generated for the data. As can be seen, the error was again only in a narrow range (−0.00091 to 0.000402).

4. Conclusions

As expected, TG and DTG showed the same shapes and trends for different heating rates, with only one main reaction. To obtain the kinetic parameters, six methods, namely, four model-free (Friedman, FWO, KAS, and Starnik) and two model-fitting (Coats–Redfern and Criado) methods were used as the main approach. A highly efficient ANN model was developed as a second approach to predict the catalytic decomposition of PET.
The two approaches were implemented to model the TGA kinetics data. In the first approach, the six different methods were used to approximate the TGA data with a straight line. Values of the activation energies obtained by Friedman, FWO, KAS, and Starnik methods were 167.5, 142.75, 138.75, and 139.13 kJ/mol, respectively. The average calculated values of activation energies by Coats–Redfern and Criado methods were in the range of 104–242 kJ/mol. For future work, more heating rates are needed to fill the gap in the results. Natural polymers, such as cellulose, are also recommended to enhance the catalytic cracking of polymer plastic waste.
An ANN with 2-15-15-1 structure and tangsig–tangsig hidden layers was determined as the most promising network as it could predict the remaining weight fraction very precisely with high regression coefficient value. The recommended network was reused with new input and still showed results that were close to the experimental values with high R and very low RMSE, MAE, and MBE.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15010070/s1, Figure S1: Criado model master plots of catalytic cracking of PET at different heating rate: (a) 2 K/min, (b) 5 K/min, (c) 10 K/min, and (d) 20 K/min; Table S1: Kinetic parameters obtained by Coats-Redfern model.

Author Contributions

Methodology, I.D.; Software, I.D.; Formal analysis, I.D.; Investigation, I.D.; Resources, I.D.; Data curation, I.D.; Writing—original draft, I.D. and Z.A.; Writing—review & editing, I.D., Z.A.; Visualization, I.D.; Supervision, I.D. Project administration, I.D. All authors have read and agreed to the published version of the manuscript.

Funding

The author are grateful to the Deanship of Scientific Research at King Faisal University (Saudi Arabia) for supporting this research as part of the Research Grants Program (Old No.: AN00029, New No.: GRANT964).

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dubdub, I.; Al-Yaari, M. Pyrolysis of Low Density Polyethylene: Kinetic Study using TGA Data and ANN Prediction. Polymers 2020, 12, 891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Martín-Gullón, I.; Esperanza, M.; Font, R. Kinetic model for the pyrolysis and combustion of poly-(ethylene terephthalate) (PET). J. Anal. Appl. Pyrolysis 2001, 58–59, 635–650. [Google Scholar] [CrossRef]
  3. 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]
  4. 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]
  5. 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]
  6. 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]
  7. 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]
  8. Saha, B.; Ghoshal, A.K. Thermal degradation kinetics of poly (ethylene terephthalate) from waste soft drinks bottles. Chem. Eng. J. 2005, 111, 39–43. [Google Scholar] [CrossRef]
  9. Aguado, J.; Serrano, D.P.; San Miguel, G.; Castro, M.C.; Madrid, S. Feedstock recycling of polyethylene in a two-step thermo-catalytic reaction system. J. Anal. Appl. Pyrolysis 2007, 79, 415–423. [Google Scholar] [CrossRef]
  10. 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]
  11. Halász, J.; Vőneki, K.; Krajkó, J. Degradation OF Pure and Waste PVC and PET in the Presence of Modified Porous Catalysts. J. Eng. Ann. Fac. Eng. Hunedoara 2009, 174, 1021–1026. [Google Scholar]
  12. 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]
  13. Lee, H.W.; Kim, Y.; Lee, B.; Kim, S.; Jae, J.; Jung, S.; Kim, T.; Park, Y. Catalytic copyrolysis of torrefied cork oak and high density polyethylene over a mesoporous HY catalyst. Catal. Today 2018, 307, 301–307. [Google Scholar] [CrossRef]
  14. 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]
  15. Xiang, Z.; Liang, J.; Morgan, H.M.J.; Liu, Y.; Mao, H.; Bu, Q. Thermal behavior and kinetic study for co-pyrolysis of lignocellulosic biomass with polyethylene over Cobalt modified ZSM-5 catalyst by thermogravimetric analysis. Bioresour. Technol. 2018, 247, 804–811. [Google Scholar] [CrossRef]
  16. 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]
  17. Ali, G.; Nisar, J.; MIqbal, M.; Shah, A.; Abbas, M.; Shah, M.R.; Rashid, U.; Bhatti, I.A.; Khan, R.A.; Shah, F. Thermo-catalytic decomposition of polystyrene waste: Comparative analysis using different kinetic models. Waste Manag. Res. 2020, 38, 202–212. [Google Scholar] [CrossRef]
  18. Lai, J.; Meng, Y.; Yan, Y.; Lester, E.; Wu, T.; Pang, C.H. Catalytic pyrolysis of linear low-density polyethylene using recycled coal ash: Kinetic study and environmental evaluation. Korean J. Chem. Eng. 2021, 38, 2235–2246. [Google Scholar] [CrossRef]
  19. Yap, T.L.; Loy, A.C.M.; Chin, B.L.F.; Lim, J.Y.; Alhamzi, H.; Chai, Y.H.; Yiin, C.L.; Cheah, K.W.; Wee, M.X.J.; Lam, M.K.; et al. Synergistic effects of catalytic co-pyrolysis Chlorella vulgaris and polyethylene mixtures using artificial neuron network: Thermodynamic and empirical kinetic analyses. J. Environ. Chem. Eng. 2022, 10, 107391. [Google Scholar] [CrossRef]
  20. Kokuryo, S.; Miyake, K.; Uchida, Y.; Mizusawa, A.; Kubo, T.; Nishiyama, N. Defect engineering to boost catalytic activity of Beta zeolite on low density polyethylene cracking. Mater. Today Sustain. 2022, 17, 100098. [Google Scholar] [CrossRef]
  21. Dourari, M.; Tarchoun, A.F.; Trache, D.; Abdelaziz, A.; Bekhouche, S.; Harrat, A.; Boukeciat, H.; Matmat, N. Unraveling the Effect of MgAl/CuO Nanothermite on the Characteristics and Thermo-Catalytic Decomposition of Nanoenergetic Formulation Based on Nanostructured Nitrocellulose and Hydrazinium Nitro-Triazolone. Catalysts 2022, 12, 1573. [Google Scholar] [CrossRef]
  22. Tarchoun, A.F.; Trache, D.; Abdelaziz, A.; Harrat, A.; Boukecha, W.O.; Hamouche, M.A.; Boukeciat, H.; Dourari, M. Elaboration, Characterization and Thermal Decomposition Kinetics of New Nanoenergetic Composite Based on Hydrazine 3-Nitro-1,2,4-triazol-5-one and Nanostructured Cellulose Nitrate. Molecules 2022, 27, 6945. [Google Scholar] [CrossRef] [PubMed]
  23. Dubdub, I.; Al-Yaari, M. Thermal Behavior of Mixed Plastics at Different Heating Rates: I. Pyrolysis Kinetics. Polymers 2021, 13, 3413. [Google Scholar] [CrossRef] [PubMed]
  24. Chan, J.H.; Balke, S.T. The thermal degradation kinetics of polypropylene: Part III. thermogravimetric analyses. Polym. Degrad. Stabil. 1997, 57, 135–149. [Google Scholar] [CrossRef]
  25. Xu, F.; Wang, B.; Yang, D.; Hao, J.; Qiao, Y.; Tian, Y. Thermal degradation of typical plastics under high heating rate conditions by TG-FTIR: Pyrolysis behaviors and kinetic analysis. Energy Convers. Manag. 2018, 171, 1106–1115. [Google Scholar] [CrossRef]
  26. Dubdub, I.; Al-Yaari, M. Pyrolysis of Mixed Waste Polymers: I: Kinetics Study. Materials 2020, 13, 4912. [Google Scholar] [CrossRef]
  27. Muravyev, N.V.; Luciano, G.; Ornaghi, H.L.J.; Svoboda, R.; Vyazovkin, S. Artificial Neural Networks for Pyrolysis, Thermal Analysis, and Thermokinetic Studies: The Status Quo. Molecules 2021, 26, 3727. [Google Scholar] [CrossRef]
  28. Quantrille, T.E.; Liu, Y.A. Artificial Intelligence in Chemical Engineering; Elsevier Science: Amsterdam, The Netherlands, 1992. [Google Scholar]
  29. Al-Yaari, M.; Dubdub, I. Application of Artificial Neural Networks to Predict the Catalytic Pyrolysis of HDPE Using Non-Isothermal TGA Data. Polymers 2020, 12, 1813. [Google Scholar] [CrossRef]
  30. Dubdub, I.; Al-Yaari, M. Pyrolysis of High Density Polyethylene: II. Artificial Neural Networks Modeling. In Proceedings of the 9th Jordan International Chemical Engineering Conference (JICHEC9), Amman, Jordan, 12–14 October 2021. [Google Scholar]
  31. Dubdub, I.; Al-Yaari, M. Pyrolysis of Mixed Plastic Waste: II. Artificial Neural Networks Prediction and Sensitivity Analysis. Appl. Sci. 2021, 11, 8456. [Google Scholar] [CrossRef]
  32. Bar, N.; Bandyopadhyay, T.K.; Biswas, M.N.; Das, S.K. Prediction of pressure drop using artificial neural network for non-Newtonian liquid flow through piping components. J. Pet. Sci. Eng. 2010, 71, 187. [Google Scholar] [CrossRef]
  33. Osman, A.I.; Fawzy, S.; Farrell, C.; Al-Muhtaseb, A.H.; Harrison, J.; Al-Mawali, S.; Rooney, D.W. Comprehensive thermokinetic modelling and predictions of cellulose decomposition in isothermal, non-isothermal, and stepwise heating modes. J. Anal. Appl. Pyrolysis 2022, 161, 105427. [Google Scholar] [CrossRef]
  34. Farrell, C.; Osman, A.I.; Harrison, J.; Vennard, A.; Murphy, A.; Doherty, R.; Russell, M.; Kumaravel, V.; Al-Muhtaseb, A.H.; Zhang, X.; et al. Pyrolysis Kinetic Modeling of a Poly(ethylene-co-vinyl acetate) Encapsulant Found in Waste Photovoltaic Modules. Ind. Eng. Chem. Res. 2021, 60, 13492–13504. [Google Scholar] [CrossRef]
  35. Muravyev, N.V.; Vyazovkin, S. The Status of Pyrolysis Kinetics Studies by Thermal Analysis: Quality Is Not as Good as It Should and Can Readily Be. Thermo 2022, 2, 435–452. [Google Scholar] [CrossRef]
  36. Koga, N.; Vyazovkin, S.; Burnham, A.K.; Favergeon, L.; Muravyev, N.V.; Pérez-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]
  37. Osman, A.I.; Young, T.J.; Farrell, C.; Harrison, J.; Al-Muhtaseb, A.H.; Rooney, D.W. Physicochemical Characterization and Kinetic Modeling Concerning Combustion of Waste Berry Pomace. ACS Sustain. Chem. Eng. 2020, 8, 17573–17586. [Google Scholar] [CrossRef]
  38. Osman, A.I.; Abdelkader, A.; Johnston, C.R.; Morgan, K.; Rooney, D.W. Thermal Investigation and Kinetic Modeling of Lignocellulosic Biomass Combustion for Energy Production and Other Applications. Ind. Eng. Chem. Res. 2017, 56, 12119–12130. [Google Scholar] [CrossRef] [Green Version]
  39. Vyazovkin, S.; Wight, C.A. Model-free and model-fitting approaches to kinetic analysis of isothermal and nonisothermal data. Thermochim. Acta 1999, 340–341, 53–68. [Google Scholar] [CrossRef]
  40. Özsin, G.; Pütün, A.E. TGA/MS/FT-IR study for kinetic evaluation and evolved gas analysis of a biomass/PVC co-pyrolysis process. Energy Convers. Manag. 2019, 182, 143–153. [Google Scholar] [CrossRef]
  41. Al-Yaari, M.; Dubdub, I. Pyrolysis of High Density Polyethylene: I. Kinetic Study. In Proceedings of the 9th Jordan International Chemical Engineering Conference (JICHEC9), Amman, Jordan, 12–14 October 2021. [Google Scholar]
  42. Al-Yaari, M.; Dubdub, I. Pyrolytic Behavior of Polyvinyl Chloride: Kinetics, Mechanisms, Thermodynamics, and Artificial Neural Network Application. Polymers 2021, 13, 4359. [Google Scholar] [CrossRef]
  43. Al-Wahaibi, T.; Mjalli, F.S. Prediction of Horizontal Oil-Water Flow Pressure Gradient Using Artificial Intelligence Techniques. Chem. Eng. Commun. 2014, 201, 209. [Google Scholar] [CrossRef]
  44. Osman, E.A.; Aggour, M.A. Artificial Neural Network Model for Accurate Prediction of Pressure Drop in Horizontal and Near-Horizontal-Multiphase Flow. Pet. Sci. Technol. 2002, 20, 1–15. [Google Scholar] [CrossRef]
  45. Qinghua, W.; Honglan, Z.; Wei, L.; Junzheng, Y.; Xiaohong, W.; Yan, W. Experimental Study of Horizontal Gas-liquid Two-phase Flow in Two Medium-diameter Pipes and Prediction of Pressure Drop through BP Neural Networks. Int. J. Fluid Mach. Syst. 2018, 11, 255–264. [Google Scholar] [CrossRef]
  46. Beale, M.; Hagan, T.; Demuth, H.B. Neural Network Toolbox TM User’s Guide; MathWorks: Natick, MA, USA, 2018. [Google Scholar]
Figure 1. Thermogravimetric analysis (TGA) and derivative thermogravimetric (DTG) curves for catalytic cracking of PET at different heating rates.
Figure 1. Thermogravimetric analysis (TGA) and derivative thermogravimetric (DTG) curves for catalytic cracking of PET at different heating rates.
Polymers 15 00070 g001
Figure 2. Linear regression lines of catalytic cracking of PET at different conversions by (a) Freidman, (b) FWO, (c) KAS, and (d) Starnik methods.
Figure 2. Linear regression lines of catalytic cracking of PET at different conversions by (a) Freidman, (b) FWO, (c) KAS, and (d) Starnik methods.
Polymers 15 00070 g002
Figure 3. Topology of the best selected (2-15-15-1) network.
Figure 3. Topology of the best selected (2-15-15-1) network.
Polymers 15 00070 g003
Figure 4. Regression of training, validation, and test plots for the selected (2-15-15-1) model.
Figure 4. Regression of training, validation, and test plots for the selected (2-15-15-1) model.
Polymers 15 00070 g004
Figure 5. Regression of simulated data for the selected (2-15-15-1) model.
Figure 5. Regression of simulated data for the selected (2-15-15-1) model.
Polymers 15 00070 g005
Figure 6. Error histogram of simulation data.
Figure 6. Error histogram of simulation data.
Polymers 15 00070 g006
Table 1. The two main analyses of PET.
Table 1. The two main analyses of PET.
PlasticProximate Analysis, wt%Ultimate Analysis, wt%
MoistureVolatileAshCHNO
PET0.52388.23111.24664.2564.367031.377
Table 2. The main properties of Zeolite β.
Table 2. The main properties of Zeolite β.
S.A. m2/gSiO2:Al2O3Pore Volume 1.7–300 nm, m2/g
68025:10.127
Table 3. Comparison between catalytic cracking characteristics of PET at different heating rates (this work) and pyrolysis of pure PET [5,6,16].
Table 3. Comparison between catalytic cracking characteristics of PET at different heating rates (this work) and pyrolysis of pure PET [5,6,16].
Heating Rate K/minMain Reaction (Catalytic Cracking)Pyrolysis of Pure PETReferences
Onset (K)Peak
(K)
Final
(K)
Mass Loss (%)Onset (K)Peak
(K)
Final
(K)
Mass Loss (%)
257567570070623667694100Osman et al. (2020) [16]
56006857107565870072380Das and Tiwari (2019) [5]
1062571074078671 (643)711 (714)748 (775)80Das and Tiwari (2019) [5], Yang et al. (2001) [6]
206507207508568172175980Das and Tiwari (2019) [5]
Table 4. Activation energies by Friedman, FWO, KAS, and Starnik methods.
Table 4. Activation energies by Friedman, FWO, KAS, and Starnik methods.
ConversionFriedmanFWOKASStarnik
E (kJ/mol)R2E (kJ/mol)R2E (kJ/mol)R2E (kJ/mol)R2
0.11300.9621970.9286910.9116920.9124
0.21580.92711190.92781150.91421150.9148
0.31780.90651370.91841320.90521330.9058
0.41850.84931490.8981450.88331450.844
0.51860.80411560.87411530.85731530.8581
0.61880.80341620.85321580.83471590.8355
0.71790.75741640.82981610.80921610.8101
0.81360.54341580.79051550.76541550.7664
Average167.50.8192142.750.8775138.750.8601139.130.8559
Table 5. Activation energy of PET catalytic pyrolysis at different heating rates by Coats–Redfern method.
Table 5. Activation energy of PET catalytic pyrolysis at different heating rates by Coats–Redfern method.
Heating Rate (K/min)E (kJ/mol)Mechanism
2146F1 (first-order chemical reaction)
104R2 (two-dimensional phase interfacial reaction)
5179F1 (first-order chemical reaction)
141R2 (two-dimensional phase interfacial reaction)
10139F1 (first-order chemical reaction)
123R2 (two-dimensional phase interfacial reaction)
20242F1 (first-order chemical reaction)
115A2 (two-dimensional nucleation and growth reaction)
224R2 (two-dimensional phase interfacial reaction)
Table 6. Distribution of 983 datasets (first step: training set) and 12 dataset (second step: simulation set) for heating rates of 2, 5, 10, 20 K/min.
Table 6. Distribution of 983 datasets (first step: training set) and 12 dataset (second step: simulation set) for heating rates of 2, 5, 10, 20 K/min.
Heating Rate (K/min)Training Set No.Simulation set No.
TrainingValidationTest
22483
52483
102463
202413
Total98312
Table 7. ANN-predicted remaining weight values of the new (simulation step) input.
Table 7. ANN-predicted remaining weight values of the new (simulation step) input.
No.Input DataOutput Data
Heating Rate
(K/min)
Temperature (K)Remaining Weight (Fraction)
12702.10.30153
22682.20.46976
32618.90.90445
45704.80.25160
55670.40.62158
65641.30.85232
710968.50.15028
810687.20.69914
910700.40.55610
1020718.60.43629
1120697.80.79087
1220732.40.20794
Table 8. Statistical parameters of the best (2-15-15-1) model.
Table 8. Statistical parameters of the best (2-15-15-1) model.
SetStatistical Parameters
R2RMSEMAEMBE
Training1.0001.820 × 10−49.800 × 10−5−4.505079 × 10−7
Validation1.0003.680 × 10−41.680 × 10−4−0.0000295
Test1.0002.580 × 10−41.390 × 10−48.015 × 10−6
All1.0002.310 × 10−41.140 × 10−4−0.00000353
Table 9. Statistical parameters of the (2-10-10-1) model for the simulated data.
Table 9. Statistical parameters of the (2-10-10-1) model for the simulated data.
SetStatistical Parameters
R2RMSEMAEMBE
simulated1.0003.903 × 10−42.843 × 10−46.360 × 10−5
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

Dubdub, I.; Alhulaybi, Z. Catalytic Pyrolysis of PET Polymer Using Nonisothermal Thermogravimetric Analysis Data: Kinetics and Artificial Neural Networks Studies. Polymers 2023, 15, 70. https://doi.org/10.3390/polym15010070

AMA Style

Dubdub I, Alhulaybi Z. Catalytic Pyrolysis of PET Polymer Using Nonisothermal Thermogravimetric Analysis Data: Kinetics and Artificial Neural Networks Studies. Polymers. 2023; 15(1):70. https://doi.org/10.3390/polym15010070

Chicago/Turabian Style

Dubdub, Ibrahim, and Zaid Alhulaybi. 2023. "Catalytic Pyrolysis of PET Polymer Using Nonisothermal Thermogravimetric Analysis Data: Kinetics and Artificial Neural Networks Studies" Polymers 15, no. 1: 70. https://doi.org/10.3390/polym15010070

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