Next Article in Journal
Influence of Binder Composition and Material Extrusion (MEX) Parameters on the 3D Printing of Highly Filled Copper Feedstocks
Next Article in Special Issue
The Effect of Various Polyhedral Oligomeric Silsesquioxanes on Viscoelastic, Thermal Properties and Crystallization of Poly(ε-caprolactone) Nanocomposites
Previous Article in Journal
Nanoparticles for Biomedical Application and Their Synthesis
Previous Article in Special Issue
Integrated Approach to Eco-Friendly Thermoplastic Composites Based on Chemically Recycled PET Co-Polymers Reinforced with Treated Banana Fibres
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Heating Rate and Temperature on the Thermal Pyrolysis of Expanded Polystyrene Post-Industrial Waste

by
Arantxa M. Gonzalez-Aguilar
,
Victoria P. Cabrera-Madera
,
James R. Vera-Rozo
and
José M. Riesco-Ávila
*
Engineering Division Mechanical, Engineering Department, University of Guanajuato, Campus Irapuato-Salamanca, Guanajuato 37320, Mexico
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(22), 4957; https://doi.org/10.3390/polym14224957
Submission received: 26 October 2022 / Revised: 9 November 2022 / Accepted: 11 November 2022 / Published: 16 November 2022
(This article belongs to the Special Issue Advanced Recycling of Plastic Waste: An Approach for Circular Economy)

Abstract

:
The use of plastic as material in various applications has been essential in the evolution of the technology industry and human society since 1950. Therefore, their production and waste generation are high due to population growth. Pyrolysis is an effective recycling method for treating plastic waste because it can recover valuable products for the chemical and petrochemical industry. This work addresses the thermal pyrolysis of expanded polystyrene (EPS) post-industrial waste in a semi-batch reactor. The influence of reaction temperature (350–500 °C) and heating rate (4–40 °C min−1) on the liquid conversion yields and physicochemical properties was studied based on a multilevel factorial statistical analysis. In addition, the analysis of the obtaining of mono-aromatics such as styrene, toluene, benzene, ethylbenzene, and α-methyl styrene was performed. Hydrocarbon liquid yields of 76.5–93% were achieved at reaction temperatures between 350 and 450 °C, respectively. Styrene yields reached up to 72% at 450 °C and a heating rate of 25 °C min−1. Finally, the potential application of the products obtained is discussed by proposing the minimization of EPS waste via pyrolysis.

1. Introduction

Plastics have played a crucial role and have been essential in the evolution of human society for 50 years because they are versatile, light, flexible, and moldable, and their production cost is low. Plastics have promoted the development of numerous applications in the automotive industry, electronics, construction, medicine, and others [1,2].
By 2020, the global production of plastics was estimated at 367 million metric tons [3,4], of which North America (United States, Mexico, and Canada) produces 19%, whereas Mexico individually contributes 2% of the global total [4,5], corresponding to 7.3 million tons approximately, which is annually increasing.
If the production and management of waste are not controlled, approximately 12,000 metric tons of plastic waste will contaminate the environment by the year 2050 [2].
Polystyrene (PS) is heat resistant, lighter in weight, and has good strength and durability, making this polymer suitable for various applications [6]. PS applications include food packaging, beverages, household appliances, the automotive field, and insulating systems for the construction industry [7,8,9,10]. Therefore, the production of PS occupies the fourth place after polyethylene (PE), polypropylene (PP), and polyvinyl chloride (PVC) [11]. The global production of expanded polystyrene (EPS) was 1.7 million tons in 2016, and the world demand for this material is constantly increasing [12]. Recycling and reusing plastic waste are essential for several reasons, the most important being conserving natural resources and reducing environmental pollution. In terms of recycling, Mexico annually consumes 125,000 tons of EPS but only recovers 1% for new-product reuse [13]. In such a scenario, there is great potential for developing an EPS recycling industry.
Plastic recycling technologies began to develop in the 1970s, and since then there have been many advances; plastic recycling can be grouped into four categories: primary, secondary, tertiary, and quaternary [14,15,16]. However, recycling plastics is difficult due to the treatments and processes required to obtain recycled products of acceptable quality. The efficiency of mechanical recycling depends on the residual plastic’s quality and the sorting process’s efficiency (primary and secondary). Materials that cannot be recycled by the mechanical method must be incinerated (quaternary) or landfilled. The high degradation stability and low density of PS cause significant problems when disposed of in landfills; therefore, the processing and recycling of PS waste is a significant problem [17].
Various technologies involving chemical recycling (tertiary) have been researched and developed including depolymerization, pyrolysis, gasification, and hydrocracking [18]. Pyrolysis is a chemical recycling technique that thermally degrades long polymer chains into small molecules in an inert environment or with limited oxygen at high temperatures (300–900 °C) [6]. The pyrolysis process generates three main products: a liquid fraction that can be used as a fuel or that can be processed into value-added chemicals; a gas fraction, which can be used to supplement the energy requirements of the process itself; and finally, a solid carbon by-product that can be used as an energy source due to its relative energy content [6,19,20,21,22]. Pyrolysis is considered an attractive method for recycling plastic waste since it transforms plastics directly into usable energy and valuable products for the chemical industry. On the other hand, the pyrolysis process is friendly to the environment because it does not produce residues [23,24]. However, not all studies have determined the composition or applications of the three resulting products.
The liquid hydrocarbon obtained by pyrolysis is not a standardized product; therefore, there are no official test methods. The liquid hydrocarbon composition varies significantly with the composition of the raw material, making it difficult to develop standards or standardized methods. Even with these challenges, efforts have been made to formalize methods for testing products to achieve a standard outcome.
Styrene is the primary aromatic compound found in the pyrolysis of EPS [17]. Benzene is another compound used to produce cyclohexane and phenol [25]. Toluene is generally used as a gasoline mixture to promote high octane, as a solvent paint, and as a precursor to synthesize other chemical compounds [12]. Additionally, ethylbenzene is widely used to generate styrene and as a solvent [26]. Xylenes and α-methyl styrene are other valuable raw materials for manufacturing plasticizers and resins [27]. In this context, the pyrolysis of EPS is an effective method for obtaining valuable chemical products in the chemical and petrochemical industry [28,29]. Therefore, recovering those chemical compounds from plastic waste is essential to the recycling industry.
As mentioned, the main objective of this study is to evaluate the influence of the reaction temperature and the heating rate on the thermal pyrolysis process of EPS waste in the conversion yields of liquid hydrocarbon, gases, and solid fractions. In addition, the effect of these input variables is compared with the aromatic compounds of the liquid products. Furthermore, the possible application of all the by-products obtained is discussed to propose a process that generates minimal waste. Moreover, the depolymerization mechanism is discussed, and the structure–property–processing relationship is addressed. Therefore, the thermal pyrolysis process of EPS waste is analyzed as a potential for developing a recycling industry in Mexico.

PS Waste Thermal Pyrolysis Overview

PS is a synthetic aromatic polymer in solid or foam form made from styrene monomer. Standard EPS comprises 98% of air and only 2% of PS [30]. Additionally, EPS waste has a chemically inert behavior, which means that it does not decompose, degrade, or disappear in the environment quickly [31].
Some of the physicochemical properties of PS are listed in Table 1. It is observed that polystyrene is mainly composed of volatile matter with a maximum of 99.59 wt.%. It also has a high carbon content (91 wt.%) and a low hydrogen and oxygen content. PS has a high calorific value; however, its use as a fuel is questionable due to its major aromatic content.
Among the parameters that have the most influence on the pyrolysis process is temperature, since it is the one that controls the cracking reaction of the polymer chain [38,39]. Some studies reported that pyrolysis at low temperatures enhances the formation of the liquid phase and the production of long hydrocarbon chains [38], while others affirm that by increasing the temperature, the yield of liquid hydrocarbon decreases, and the formation of gases increases. High temperatures improve the secondary reactions inside the reactor, which reduce the obtained solid by-product [40].
In terms of state of the art on thermal pyrolysis of PS, the work of Lu et al. [33] is highlighted, where they experimented with virgin PS in a reactor under an inert atmosphere with nitrogen. The process was carried out at a heating rate of 5 °C min−1 up to a reaction temperature of 420 °C for 120 min. Their results reported a liquid hydrocarbon yield of 76.26%, and 73% styrene stands out in its composition.
On the other hand, Verma et al. [12] evaluated the influence of temperature (400–700 °C) and heating rate (5–25 °C min−1) of thermal pyrolysis of polystyrene waste. The pyrolysis process was carried out in a batch-type reactor and under an inert atmosphere with nitrogen at a flow of 200 mL min−1. Regarding the heating rate, the results concluded that the liquid yield increased as the heating rate increased; however, higher rates decreased its yield. The optimum temperature found under these conditions was 650 °C and a heating rate of 15 °C min−1, obtaining 94.37% of liquid hydrocarbon. The composition of the liquid hydrocarbon corresponded to a concentration of 84.74% of styrene.
A recent study developed by Van der Westhuizen et al. [35] evaluated the thermal pyrolysis process of three different types of PS to analyze the effect of contamination of the raw material on fuel production. The research studied, as input factors, the temperature and the heating rate in a semi-batch reactor. Additionally, the study industrially scaled the process to a semi-continuous rotary reactor and analyzed the properties of the fuel. The results indicated that contamination in the polystyrene raw material can decrease the liquid yield by up to 6.4%; however, it did not significantly affect its energy content. Van der Westhuizen et al. highlighted the importance of blending PS pyrolytic oil with some other transportation fuel due to its high aromatic content.

2. Materials and Methods

2.1. Materials

The raw material used in this work was EPS waste, generated during packaging manufacturing. This waste was supplied by a company in Irapuato, Guanajuato, Mexico. Analytical grade reagents (toluene, benzene, ethylbenzene, xylenes, and styrene) were purchased from Sigma-Aldrich.

2.2. Thermogravimetric Analysis (TGA)

Thermogravimetric analysis is commonly used to consider the degradation trend in terms of different parameters of the pyrolysis process, such as temperature, heating rate, and others [41]. TGA can be used mainly to study the degradation behavior of polymeric materials, including homopolymers, copolymers, and others [42]. This study performed the thermogravimetric analysis of EPS using a TA Instruments SDT Q600 thermobalance (New Castle, DE, USA). The initial mass of the sample was 4.41 mg. The experiment was carried out under an inert atmosphere with nitrogen gas (N2 5.0) at a 20 mL min−1 rate and a heating rate of 20 °C min−1 up to 600 °C.

2.3. Pyrolytic Oil Characterization

One of the main physical properties of a material is density; in this study, the density was evaluated by buoyancy using a glass hydrometer, and the measurement was performed under the ASTM D 1298 standard, at 20 °C. For this study, a Cannon-Fenske viscometer was used, and the measurement of kinematic viscosity was performed under the ASTM D445 standard at 40 °C. On the other hand, the heating value was determined using an IKA C3000®® isoperibolic bomb calorimeter (Staufen, Germany) under the ASTM D4809 standard.

2.4. Gas Chromatography

The qualitative analysis and identification of the chemical compounds present were carried out using a Varian®® 450 GC gas chromatograph (Waltham, MA, USA). The GC was equipped with an Omegawax®® 250 fused silica capillary column, 30 m × 0.25 m × 0.25 μm, using benzene, cumene, styrene, ethylbenzene, and toluene as standards. Helium was used as the carrier gas at a 25 mL min−1 flow rate. The injection volume was 1 µL with a split ratio of 1:50. The injection temperature was 250 °C. For temperature programming, the oven was held at a temperature of 40 °C for one minute and then increased to 200 °C at a 10 °C min−1 rate; it was then increased to 240 °C at a rate of 5 °C min−1 and maintained for 15 min. For the quantitative analysis of the products, C19 was used as the internal standard.

2.5. Pyrolysis Experimental Setup

This study carried out the thermal pyrolysis of EPS waste in a semi-batch type reactor. The reactor consisted of a stainless-steel tube 17 cm high and 4.5 inches in nominal diameter. The reactor was heated by an external band-type electrical resistance with a ceramic core. Temperatures were monitored using k-type thermocouples inside the reactor and controlled by PID temperature control within ±4 °C.
The experimental scheme of the studied experimental process is shown in Figure 1. The reactor was fed with EPS waste and was hermetically sealed to prevent leaks without the addition of any solvent or any inert gas. The reactor was heated, and the experienced temperatures considered in this study were those obtained through the thermocouple (TC1). The gases produced by thermal pyrolysis were dislodged through a pipe that flowed into a distiller manufactured according to the ASTM D86 standard; this consisted of a pipe submerged in a container with water at room temperature. The gases that were not condensed in the first stage passed through a second countercurrent flow condenser. The condensed product was stored in the secondary collector, the gases that were not condensed in this stage passed to a water trap unit, and, finally, the non-condensable gases were released. The solid fraction was collected from the bottom of the reactor.
Pyrolysis residence time started when the temperature measured by the thermocouple (TC2) reached 150 °C and ended 30 min after the thermocouple (TC1) reached the desired temperature.

2.6. Operation Parameters

Statistical analysis was performed with the experimental data obtained in the present study. The analysis was performed using Statgraphics Centurion software XVI with a multilevel factorial design, which is used to study effects with n quantitative factors. The input variables were the heating rate and the reaction temperature; in contrast, the output responses focused on the percentage of conversion to liquid hydrocarbon and the formation of styrene. Both input variables had four levels, and a replica was made for each experiment representing 32 runs.
The yields of liquid, solid, and gas fractions were calculated using Equations (1)–(3), respectively:
Liquid   yield   ( wt . % ) = ( Liquid   mass / EPS   mass )   ×   100
Solid   yield   ( wt . % ) = ( Solid   mass / EPS   mass )   ×   100
Gas   yield   ( wt . % ) = [ ( Liquid   mass + Solid   mass / EPS   mass ) ]   ×   100

3. Results

3.1. Thermogravimetric Analysis

Figure 2 shows the calorimetric curve indicating the non-isothermal mass loss of EPS measured by a thermogravimetric analyzer (TGA) at a heating rate of 20 °C min−1. The results show that the sample tended to degrade at temperatures higher than 350 °C, obtaining a 98% mass loss at 454 °C, approximately. The final degradation temperature was close to what was reported by Kremer et al. [43], which was 448 °C for PS at the same rate of 20 °C min−1. Fuentes et al. [44] reported temperatures close to 458 °C for virgin PS and PS wastes.
As noted, these slight differences between the degradation temperatures of the EPS sample used for the TGA in this work and the ones reported by Kremer et al. [43] and Fuentes et al. [44] were the product of several factors, such as the preparation method of polymer, particle size, the molecular weight of the polymer, operating conditions of the thermogravimetric apparatus, and the mathematical treatment of thermogravimetric data [45,46,47].
Even when those factors influenced the degradation temperatures, activation energy, and kinetic behavior, they did not affect the overall thermal decomposition. The TGA/DTG plots (Figure 2) for the thermal decomposition of the EPS revealed a one-step degradation in the temperature range of 350 °C to 450 °C, which is consistent with the data reported in the literature [46,47,48,49,50,51].
EPS has shown a glass transition temperature of around 100 °C [52], and when thermally decomposed melts at about 160 °C, and the volatility of molten polymer with high molecular weight decreases at 275 °C [47]. This behavior could be seen in the first stage of the exothermic reaction (blue line) when it surpassed 288.20 °C and reached 375.81 °C.
At this point, the single-phase degradation occurred (green line), and it was attributed to the decomposition of the EPS solid matrix to volatile styrene monomers and derivatives (fragments with low molecular weight), reaching 431.66 °C to 454 °C for the complete degradation reported by Mehta et al. [48] and Ali et al. [47]. It is noteworthy to mention that this weight loss indicated that thermal degradation of EPS in a non-inert atmosphere will show a reductive behavior.

3.2. Influence of Temperature on the Performance of the EPS Pyrolysis Process

Table 2 shows the conversion yields of the products obtained at different pyrolysis temperatures grouped by the heating rate experimented. It was observed that the production of liquid hydrocarbon increased as the pyrolysis temperature increased. For any heating rate, the maximum liquid hydrocarbon yield was obtained at 450 °C, which was consistent with results from TGA. In this context, with temperatures above 450 °C, the liquid hydrocarbon yield began to decrease by a maximum of 3 wt.%, related to the reduction in the fragments of high molecular weight and the increment of volatile styrene monomers and derivatives (low molecular weight), as well other gas products of the reductive atmosphere.
Additionally, Maafa [6] reported that if the preferred product in the PS pyrolysis process is liquid, it is recommended to use a temperature range of 350 to 500 °C. In contrast, if ash or gases are desired as a product, temperatures above 500 °C are indicated. The yield of gases and solid fractions is significant at a temperature of 350 °C, while there is no significant difference for the other temperatures. This is explained by the fact that in the range of 350 °C and 450 °C, the single-phase thermal degradation of EPS is a radical chain process characterized by three consecutive steps: (1) initiation, (2) propagation, and (3) termination [47,53,54].
During this time, the diffusion of heat or decomposition gases has to be considered as a simultaneous process to the overall chemical reaction (including intermolecular condensation reactions), which has an endothermal/exothermal behavior, inducing heterogeneous temperature distribution in the reactor with no effect on the general composition [55,56].
Particularly in semi-batch systems, like the one used in this study, the evaporated volatiles are removed from the heated zone by evacuation or purge and sweep gases [57] Consequently, thermal degradation occurs only in the liquid phase, and the remnant cracking reactions in the gas phase are negligible [58].
The results of this study were similar to those reported by Tamri et al. [41], who studied the high-impact polystyrene (HIPS) pyrolysis process and obtained maximum yields at 450 °C. In the present study, a maximum yield of 91 wt.% was obtained, having a difference of 3.9%, respective to their study, but without the two additional parameters used by them: an inert environment with nitrogen gas flow and a stirring of 50 rpm. Even compared to catalytic pyrolysis processes, which work at higher temperatures (superior to 600 °C), with the process addressed in this study, it is possible to recover higher values in liquid hydrocarbon [6,59,60,61,62].

3.3. Influence of Heating Rate in Obtaining Liquid Hydrocarbon

The heating rate is a parameter that influences the pyrolysis process, directly impacting kinetic behavior; studies have observed that a higher heating rate enhances the production of ashes and gases, reducing the yield of liquid hydrocarbon [63]. Figure 3 shows the results obtained evaluating the four heating ramps (4, 12, 25, and 40 °C min−1). It was observed that for the heating rates experienced, both the pyrolysis temperature and the effect of the heating rate had quadratic responses in the yield of liquid hydrocarbon. At a higher heating rate, it will favor the production of liquid; however, it has a maximum point at a 12 °C min−1 rate. With higher values, a decrease in performance will be observed. Nanda and Berruti [7] stated that rapid pyrolysis and high degradation temperatures tend to decrease the yield of the plastic pyrolysis liquid. This is due to the faster achieving of the exothermic phase, where the weight loss is at its maximum, which, for the EPS sample used in this work, was 98% at 454 °C, as mentioned before.
Reaching the exothermic phase so fast implies that the endothermic phase, where the polymer suffers the glass transition, has a minimum time to break the larger molecules and melt into a polymer with high molecular weight. Moreover, since the thermal degradation occurs only in this liquid phase and the reactions in the gas phase are negligible [58], the volatile compounds produced are minimal. The last are those that after condensation become a pyrolytic liquid oil, which is the main product of interest in this work.

3.4. Influence of Temperature and Heating Rate in Obtaining Value-Added Products

The pyrolysis of EPS is the process where the highest conversion percentage is obtained among all plastics. However, the pyrolytic liquid cannot be used as fuel due to its aromatic composition (principally styrene and α-methyl styrene), which causes a very low thermal–oxidative stability [17] and, therefore, increased engine carbon deposition (if used as automobile fuel) [56]. Thus, the application of this product is mainly based on the petrochemical industry. Figure 4 shows the results of the compounds obtained in the temperatures experimented, grouped in the four heating rate studies. It was observed that for the thermal pyrolysis of EPS in a semi-batch reactor and regardless of the temperature, the main product was a liquid rich in aromatic compounds such as styrene, toluene, ethylbenzene, and α-methyl styrene. The results showed that styrene was the aromatic compound found in the highest proportion. The maximum styrene concentration found was 72.99% at a temperature of 450 °C and a heating rate of 25 °C min−1. In contrast, the minimum was found at 51.28% at a temperature of 350 °C and a heating rate of 40 °C min−1. This is explained by the fact that in the range of 350 to 450 °C, the single-phase thermal degradation of EPS was occurring, as has been stated before.
Toluene concentrations varied from 4.28 to 12.31% at 350 and 450 °C, respectively. In the chromatographic analysis, values lower than 0.5% were discarded for this study. Additionally, as by-products of the process, the solid fraction was considered to be mainly coal and other non-degradable residues and non-condensable gases. According to the literature, the significant components in the gaseous fraction are related to alkane and alkene. These non-condensable compounds and non-water-soluble gases are methane, ethane, ethene (ethylene), propane, and pentene [63,64,65]. The presence of nitrogen has also been reported [64].
The reaction mechanism of the EPS thermochemical degradation is a chain reaction that begins with a random scission (β-scission) that forms macroradicals (C13–C24 fraction) along with styrene [66,67,68]. The second stage, called propagation, is a series of intermolecular hydrogen transfers, initially forming low molecular weight macromolecules (C6–C11 fraction), and in the final stages, dimers and trimers are derived from styrene [55,68,69]. This is the best fitted reaction model, which involves around 2700 to 4500 reactions happening simultaneously, entailing 64 to 93 dead and live species, according to the literature [70,71,72]. Based on the above and the results of the chromatographic analysis of the present study, the proposed reaction mechanism for the degradation of polystyrene is shown in Figure 5.
Toluene, ethylbenzene, and cumene are intermediates between the fractions, donating or accepting hydrogens during the propagation stage because to degrade styrene monomers, dimers, and trimers, many hydrogen radicals are required [73]. This is correlated with the change in styrene selectivity [17], which decreases as the temperature and the heating rate increase [63], as can be seen in Figure 4, where at 500 °C a noticeable fall is shown for the heating rates of 25 and 40 °C min−1. Furthermore, as mentioned before, the thermal degradation of EPS only happens in the liquid phase and between the range of 350 to 450 °C, reaching complete degradation at 470 °C [47,48].
Ethylbenzene is a hydrogenated compound derived from styrene, which explains the relationship between both and their mirrored behavior, as shown in Figure 4. As the styrene concentration increased, the concentration of ethylbenzene decreased, and vice versa. In addition, the relationship between α-methyl styrene and cumene had an opposite behavior, because even though the latter is the hydrogenated compound of the former, the presence of cumene was only detected when the concentration of α-methyl styrene increased considerably. This corresponded to the diminution of styrene at the highest temperature (500 °C), where complete degradation of the EPS sample was reached.
On the other hand, the presence of toluene, even when it showed the same tendency as styrene for any temperature, was in a significant percentage at 500 °C, also corresponding with the lowest styrene concentration. This factor had an overall boost of benzene and cumene compounds, which were only shown at this temperature. Consequently, the obtained oil predominantly comprised styrene monomers, toluene, benzene, styrene dimers, and styrene trimers [63,64]. These oxygenated compounds are the product of the reductive atmosphere inside the reactor and the trapped oxygen in the EPS [63].

3.5. Characterization of by-Products from EPS Waste Pyrolysis

Table 3 shows the physicochemical properties of the liquid fraction with the highest liquid yield and highest styrene concentration found and is compared with the literature. It was observed that the highest liquid yields resulted at heating rates of 12 to 25 °C min−1, and low heating rates (5 °C min−1) decreased the conversion to liquid. The results of the present investigation indicated that temperatures above 500 °C decreased the conversion to a liquid fraction. In this context, it was compared with the results of Van der Westhuizen et al. [35], who reported the characterization of the PS pyrolytic oil with the highest yield of 82.5 wt.% at 550 °C, which is 8.5 wt.% lower than that obtained at 500 °C in the present investigation.
On the other hand, no significant difference in styrene formation was observed between temperatures from 420 to 450 °C. Likewise, at higher temperatures, a lower yield percentage was found. However, the objective of the work carried out by Westhuizen et al. was to reduce the styrene concentration to propose it as a fuel, in the case of the physicochemical properties, density, and viscosity decreases at temperatures close to 550 °C. However, the application to be given to the pyrolytic liquid must be considered to indicate the required values. Finally, the energy content of the pyrolytic oil derived from EPS waste is higher than 41 MJ/kg, so its use as an energy source would be advantageous. Nevertheless, when burned, the aromatic content must be considered.
An essential aspect of the pyrolysis of plastics is identifying the applications of all the by-products. Most of the literature focuses on liquid hydrocarbons; however, few have studied the possible applications of non-condensable gases, which are typically burnt [74]. However, converting non-condensable gases into valuable materials can improve the economy of waste-to-liquid production and make thermochemical processing more competitive compared to other recycling technologies.
Singh et al. [63] mentioned and Veksha et al. [75] studied the conversion of non-condensable gases from the pyrolysis of plastics, including PS, into carbon nanotubes (NTCs). NTCs show a diversity of potential applications, including photothermal therapy, photoacoustic imaging, biomedicine, and even an alternative to removing air pollutants [76].
Additionally, Miandad et al. [77] used PS pyrolysis-derived char for the synthesis of nano-absorbent carbon metal double layer oxides (C/MnCuAl-LDOs) through co-precipitation for the adsorption of Congo red (CR) from wastewater. The nano-absorbent from PS pyrolysis was compared with pure carbon (PC), thermally activated carbon (TAAC), and oxidized carbon (Ox-C). Their results reported that C/MnCuAl-LDOs showed maximum adsorption capacity for CR among all the used absorbents.
Moreover, Dogu et al. [78] mentioned that PS, when pyrolyzed with other solid plastic waste, produces a valuable liquid oil and gases that by aromatization or catalytic reforming could produce an aromatic blend with the potential to be used as aviation fuel.

3.6. Statistical Analysis

3.6.1. Liquid Hydrocarbon

Figure 6 shows the standardized Pareto chart for the conversion to liquid oil yield with a confidence interval of α = 0.5 and t = 2.05 (blue line). It was observed that the factor with the most significant influence on liquid performance was temperature, having a positive effect. As the temperature increased, the liquid yield increased. On the other hand, there was a significant influence on the interaction between temperature and heating rate.
Figure 7 shows the main effects of temperature and heating rate on the conversion to liquid hydrocarbons. The analysis was performed in a second model. The order allowed for a better approximation model for the phenomenon being studied. It was observed that the response of the variables had a curved effect; as the temperature increased, the liquid yield increased. However, temperatures above 470 °C would decrease the conversion of liquid hydrocarbon. On the other hand, the curved effect of the heating rate on the output response was also observed; heating rates greater than 4 °C min−1 but less than 12 °C min−1 would improve liquid hydrocarbon production.
Additionally, Equation (4) describes the fitted multiple regression model (R2 = 80.40%) of the conversion to liquid hydrocarbon, where T is the reaction temperature, and H is the heating rate.
% L i q u i d = 67.3048 + 0.649545 × T + 0.805674 × H 0.0006625 × T 2 0.00134173 × T × H 0.0066103 × H 2

3.6.2. Styrene Production

This section shows the statistical analysis results of styrene production as an output response, evaluating the reaction temperature and the heating rate as input variables. Figure 8 shows the standardized Pareto chart for styrene production by an analysis with a confidence interval of α = 0.5 and t = 2.05 (blue line). It was observed that the factor with the most significant influence on the formation of styrene was the heating rate. In addition, it was visualized that this effect was negative, representing that at a higher heating rate, less styrene will be formed. On the other hand, the reaction temperature also influenced the output response; however, its effect was positive. At a higher reaction temperature, the liquid hydrocarbon composition would have a higher concentration of styrene. There was no interaction between both factors that influenced the formation of styrene.
In the same way as the conversion to liquid hydrocarbon, the main effects of the input variables on the production of styrene were analyzed. It is visualized in Figure 9 that both temperature and heating rate had curved effects on the output responses. It is crucial to analyze nonlinear models since sometimes the linear trend makes false statements. Therefore, it was statistically stated that there was a decrease in styrene production at temperatures greater than 450 °C and at heating rates above 17 °C min−1.
Regarding the multiple regression model that describes the formation of styrene, Equation (5) describes the behavior with a fit of 77%, where T is the reaction temperature, and H is the heating rate.
% S t y r e n e = 91.8019 + 0.65818 × T + 1.43878 × H 0.000694625 × T 2 0.0013590 × T × H 0.0239486 × H 2
Finally, in Figure 10, the response surface of the styrene production is shown through a contour plot. It is seen that the highest concentration of styrene in liquid hydrocarbons obtained by thermal pyrolysis of EPS was centered at temperatures below 500 °C and maximum heating rates of 20 °C min−1.

4. Conclusions

The influence of temperature and heating rate in the thermal pyrolysis process of (EPS) post-industrial waste shows an intrinsic relationship between the production of liquid hydrocarbon yield and styrene content. Moreover, neither the temperature nor the heating rate showed an influence on the general composition found in the pyrolytic oil.
The results showed a maximum yield of liquid hydrocarbon of 91 wt.%, with 72.99% of styrene in its composition, at 450 °C and a heating rate of 25 °C min−1. It is noteworthy that in a simple thermal pyrolysis process as the one addressed in this study, it was possible to recover higher values in liquid hydrocarbon compared to other studies. It was even compared to catalytic pyrolysis processes, which achieve these yields at higher temperatures (superior to 600 °C) and by adding elements that lead to higher operational costs [6,59,60,61,62].
This proposed pyrolysis can go hand-in-hand with EPS mechanical recycling, which will solve the limitations of this type of plastic waste management but also for producing a liquid hydrocarbon with added value for the petrochemical industry.

Author Contributions

Conceptualization, methodology, and experimental design, A.M.G.-A.; experimentation and formal analysis, A.M.G.-A., V.P.C.-M. and J.R.V.-R.; writing—original draft preparation, A.M.G.-A. and V.P.C.-M.; writing—review and editing, J.M.R.-Á. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by UNIVERSIDAD DE GUANAJUATO, grant number 058/2022 CIIC.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the University of Guanajuato for sponsorship of this paper and the “Poliespuma del Bajío” company for supplying the raw material for this investigation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jeswani, H.; Krüger, C.; Russ, M.; Horlacher, M.; Antony, F.; Hann, S.; Azapagic, A. Life cycle environmental impacts of chemical recycling via pyrolysis of mixed plastic waste in comparison with mechanical recycling and energy recovery. Sci. Total Environ. 2021, 769, 144483. [Google Scholar] [CrossRef]
  2. Geyer, R.; Jambeck, J.R.; Law, K.L. Production, use, and fate of all plastics ever made. Sci. Adv. 2017, 3, e1700782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Statista online Annual Production of Plastics Worldwide from 1950 to 2020. Available online: https://www.statista.com/statistics/282732/global-production-of-plastics-since-1950/#statisticContainer (accessed on 1 August 2022).
  4. Plastics Europe World and European Plastics Production Evolution 2021. Available online: https://plasticseurope.org/wp-content/uploads/2021/12/Plastics-the-Facts-2021-web-final.pdf (accessed on 1 August 2022).
  5. Condor Ferries Plastics in the Ocean. Available online: https://www.condorferries.co.uk/plastic-in-the-ocean-statistics (accessed on 1 August 2022).
  6. Maafa, I. Pyrolysis of Polystyrene Waste: A Review. Polymers 2021, 13, 225. [Google Scholar] [CrossRef] [PubMed]
  7. Nanda, S.; Berruti, F. Thermochemical conversion of plastic waste to fuels: A review. Environ. Chem. Lett. 2020, 19, 123–148. [Google Scholar] [CrossRef]
  8. Dewangga, P.B.; Rochmadi; Purnomo, C.W. Pyrolysis of polystyrene plastic waste using bentonite catalyst. IOP Conf. Series Earth Environ. Sci. 2019, 399, 012110. [Google Scholar] [CrossRef]
  9. Amjad, U.; Ishaq, M.; Rehman, H.U.; Ahmad, N.; Sherin, L.; Hussain, M.; Mustafa, M. Diesel and gasoline like fuel production with minimum styrene content from catalytic pyrolysis of polystyrene. Environ. Prog. Sustain. Energy 2020, 40, e13493. [Google Scholar] [CrossRef]
  10. Klemencová, K.; Grycová, B.; Inayat, A.; Leštinský, P. thermo-catalytic degradation of polystyrene Over α-Fe2O3. Nanocon 2020, 267–271. [Google Scholar] [CrossRef]
  11. Inayat, A.; Klemencova, K.; Grycova, B.; Sokolova, B.; Lestinsky, P. Thermo-catalytic pyrolysis of polystyrene in batch and semi-batch reactors: A comparative study. Waste Manag. Res. J. Sustain. Circ. Econ. 2020, 39, 260–269. [Google Scholar] [CrossRef] [PubMed]
  12. Verma, A.; Sharma, S.; Pramanik, H. Pyrolysis of waste expanded polystyrene and reduction of styrene via in-situ multiphase pyrolysis of product oil for the production of fuel range hydrocarbons. Waste Manag. 2020, 120, 330–339. [Google Scholar] [CrossRef]
  13. Plastic’s Technology Mexico Builds EPS Recycling Industry. Available online: https://www.plastico.com/temas/Mexico-construye-una-industria-del-reciclaje-de-EPS+125381 (accessed on 1 August 2022).
  14. Ignatyev, I.; Thielemans, W.; Beke, B.V. Recycling of Polymers: A Review. ChemSusChem 2014, 7, 1579–1593. [Google Scholar] [CrossRef]
  15. Panda, A.K. Studies on Process Optimization for Production of Liquid Fuels from Waste Plastics. Ph.D. Thesis, National Institute of Technology Rourkela, Rourkela, India, 2011. [Google Scholar]
  16. Ragaert, K.; Delva, L.; Van Geem, K. Mechanical and chemical recycling of solid plastic waste. Waste Manag. 2017, 69, 24–58. [Google Scholar] [CrossRef] [PubMed]
  17. Dement’Ev, K.I.; Palankoev, T.A.; Alekseeva, O.A.; Babkin, I.A.; Maksimov, A.L. Thermal depolymerization of polystyrene in highly aromatic hydrocarbon medium. J. Anal. Appl. Pyrolysis 2019, 142, 104612. [Google Scholar] [CrossRef]
  18. Grigore, M.E. Methods of Recycling, Properties and Applications of Recycled Thermoplastic Polymers. Recycling 2017, 2, 24. [Google Scholar] [CrossRef] [Green Version]
  19. Veses, A.; Aznar, M.; Martínez, I.; Martínez, J.; López, J.; Navarro, M.; Callén, M.; Murillo, R.; García, T. Catalytic pyrolysis of wood biomass in an auger reactor using calcium-based catalysts. Bioresour. Technol. 2014, 162, 250–258. [Google Scholar] [CrossRef]
  20. Veses, A.; Aznar, M.; Callén, M.; Murillo, R.; García, T. An integrated process for the production of lignocellulosic biomass pyrolysis oils using calcined limestone as a heat carrier with catalytic properties. Fuel 2016, 181, 430–437. [Google Scholar] [CrossRef]
  21. Yildiz, G.; Ronsse, F.; van Duren, R.; Prins, W. Challenges in the design and operation of processes for catalytic fast pyrolysis of woody biomass. Renew. Sustain. Energy Rev. 2016, 57, 1596–1610. [Google Scholar] [CrossRef]
  22. Anuar Sharuddin, S.D.; Abnisa, F.; Wan Daud, W.M.A.; Aroua, M.K. A review on pyrolysis of plastic wastes. Energy Convers. Manag. 2016, 115, 308–326. [Google Scholar] [CrossRef]
  23. Qureshi, M.S.; Oasmaa, A.; Pihkola, H.; Deviatkin, I.; Tenhunen, A.; Mannila, J.; Minkkinen, H.; Pohjakallio, M.; Laine-Ylijoki, J. Pyrolysis of plastic waste: Opportunities and challenges. J. Anal. Appl. Pyrolysis 2020, 152, 104804. [Google Scholar] [CrossRef]
  24. Antelava, A.; Damilos, S.; Hafeez, S.; Manos, G.; Al-Salem, S.M.; Sharma, B.K.; Kohli, K.; Constantinou, A. Plastic Solid Waste (PSW) in the Context of Life Cycle Assessment (LCA) and Sustainable Management. Environ. Manag. 2019, 64, 230–244. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, J.; Jiang, J.; Sun, Y.; Zhong, Z.; Wang, X.; Xia, H.; Liu, G.; Pang, S.; Wang, K.; Li, M.; et al. Recycling benzene and ethylbenzene from in-situ catalytic fast pyrolysis of plastic wastes. Energy Convers. Manag. 2019, 200, 112088. [Google Scholar] [CrossRef]
  26. Gaurh, P.; Pramanik, H. Production of benzene/toluene/ethyl benzene/xylene (BTEX) via multiphase catalytic pyrolysis of hazardous waste polyethylene using low cost fly ash synthesized natural catalyst. Waste Manag. 2018, 77, 114–130. [Google Scholar] [CrossRef] [PubMed]
  27. Cai, Q.; Li, J.; Bao, F.; Shan, Y. Tunable dimerization of α-methylstyrene catalyzed by acidic ionic liquids. Appl. Catal. A Gen. 2005, 279, 139–143. [Google Scholar] [CrossRef]
  28. Sarker, M.; Rashid, M.M.; Molla, M.; Rahman, M.S. Thermal Conversion of Waste Plastics (HDPE, PP and PS) to Produce Mixture of Hydrocarbons. Am. J. Environ. Eng. 2012, 2, 128–136. [Google Scholar] [CrossRef]
  29. Xu, X.; Jiang, E.; Li, Z.; Sun, Y. BTX from anisole by hydrodeoxygenation and transalkylation at ambient pressure with zeolite catalysts. Fuel 2018, 221, 440–446. [Google Scholar] [CrossRef]
  30. Gil-Jasso, N.D.; Segura-González, M.A.; Soriano-Giles, G.; Neri-Hipolito, J.; López, N.; Mas-Hernández, E.; Barrera-Díaz, C.E.; Varela-Guerrero, V.; Ballesteros-Rivas, M.F. Dissolution and recovery of waste expanded polystyrene using alternative essential oils. Fuel 2018, 239, 611–616. [Google Scholar] [CrossRef]
  31. Gutierrez-Velasquez, E.I.; Monteiro, S.N.; Colorado, H.A. Characterization of expanded polystyrene waste as binder and coating material. Case Stud. Constr. Mater. 2022, 16. [Google Scholar] [CrossRef]
  32. Jaafar, Y.; Abdelouahed, L.; El Hage, R.; El Samrani, A.; Taouk, B. Pyrolysis of common plastics and their mixtures to produce valuable petroleum-like products. Polym. Degrad. Stab. 2021, 195, 109770. [Google Scholar] [CrossRef]
  33. Lu, C.; Xiao, H.; Chen, X. Simple pyrolysis of polystyrene into valuable chemicals. e-Polymers 2021, 21, 428–432. [Google Scholar] [CrossRef]
  34. Park, S.S.; Seo, D.K.; Lee, S.H.; Yu, T.-U.; Hwang, J. Study on pyrolysis characteristics of refuse plastic fuel using lab-scale tube furnace and thermogravimetric analysis reactor. J. Anal. Appl. Pyrolysis 2012, 97, 29–38. [Google Scholar] [CrossRef]
  35. van der Westhuizen, S.; Collard, F.; Görgens, J. Pyrolysis of waste polystyrene into transportation fuel: Effect of contamination on oil yield and production at pilot scale. J. Anal. Appl. Pyrolysis 2021, 161, 105407. [Google Scholar] [CrossRef]
  36. 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]
  37. Martín-Lara, M.; Piñar, A.; Ligero, A.; Blázquez, G.; Calero, M. Characterization and Use of Char Produced from Pyrolysis of Post-Consumer Mixed Plastic Waste. Water 2021, 13, 1188. [Google Scholar] [CrossRef]
  38. Klaimy, S.; Lamonier, J.-F.; Casetta, M.; Heymans, S.; Duquesne, S. Recycling of plastic waste using flash pyrolysis—Effect of mixture composition. Polym. Degrad. Stab. 2021, 187, 109540. [Google Scholar] [CrossRef]
  39. Jha, K.K.; Kannan, T. Recycling of plastic waste into fuel by pyrolysis—A review. Mater. Today Proc. 2020, 37, 3718–3720. [Google Scholar] [CrossRef]
  40. López, A.; de Marco, I.; Caballero, B.; Laresgoiti, M.; Adrados, A. Influence of time and temperature on pyrolysis of plastic wastes in a semi-batch reactor. Chem. Eng. J. 2011, 173, 62–71. [Google Scholar] [CrossRef]
  41. Tamri, Z.; Yazdi, A.V.; Nekoomanesh, M.; Abbas-Abadi, M.S.; Heidarinasab, A. The effect of temperature, heating rate and zeolite based catalysts on the pyrolysis of high impact polystyrene (HIPS) waste to produce the fuel like products. Polyolefins J. 2018, 6, 43–52. [Google Scholar] [CrossRef]
  42. Ren, X.; Huang, Z.; Wang, X.-J.; Guo, G.-M. Isoconversional analysis of kinetic pyrolysis of virgin polystyrene and its two real-world packaging wastes. J. Therm. Anal. 2021, 147, 1421–1437. [Google Scholar] [CrossRef]
  43. Kremer, I.; Tomić, T.; Katančić, Z.; Hrnjak-Murgić, Z.; Erceg, M.; Schneider, D.R. Catalytic decomposition and kinetic study of mixed plastic waste. Clean Technol. Environ. Policy 2020, 23, 811–827. [Google Scholar] [CrossRef]
  44. Fuentes, C.; Lerner, J.C.; Vázquez, P.; Sambeth, J. Analysis of the emission of PAH in the thermal and catalytic pyrolysis of polystyrene. Catal. Today 2021, 372, 175–182. [Google Scholar] [CrossRef]
  45. Carrasco, F.; Pagès, P. Thermogravimetric Analysis of Polystyrene: Influence of Sample Weight and Heating Rate on Thermal and Kinetic Parameters. J. Appl. Polym. Sci. 1996, 61, 187–197. [Google Scholar] [CrossRef]
  46. Singh, R.; Ruj, B.; Sadhukhan, A.; Gupta, P. Thermal degradation of waste plastics under non-sweeping atmosphere: Part 1: Effect of temperature, product optimization, and degradation mechanism. J. Environ. Manag. 2019, 239, 395–406. [Google Scholar] [CrossRef] [PubMed]
  47. Ali, G.; Nisar, J.; Iqbal, 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. J. Sustain. Circ. Econ. 2019, 38, 202–212. [Google Scholar] [CrossRef] [PubMed]
  48. Mehta, S.; Biederman, S.; Shivkumar, S. Thermal degradation of foamed polystyrene. J. Mater. Sci. 1995, 30, 2944–2949. [Google Scholar] [CrossRef]
  49. Zeng, W.R.; Chow, W.; Yao, B. Chemical Kinetics and Mechanism of Polystyrene Thermal Decomposition. Fire Safety Sci. 2007, 7. [Google Scholar]
  50. Hachani, S.E.; Wis, A.A.; Necira, Z.; Nebbache, N.; Meghezzi, A.; Ozkoc, G. Effects of Magnesia Incorporation on Properties of Polystyrene/Magnesia Composites. Acta Chim. Slov. 2018, 65, 646–651. [Google Scholar] [CrossRef] [PubMed]
  51. Farha, A.H.; Al Naim, A.F.; Mansour, S.A. Thermal Degradation of Polystyrene (PS) Nanocomposites Loaded with Sol Gel-Synthesized ZnO Nanorods. Polymers 2020, 12, 1935. [Google Scholar] [CrossRef] [PubMed]
  52. Luo, S.; Gao, L.; Guo, W. Effect of expanded polystyrene content and press temperature on the properties of low-density wood particleboard. Maderas. Ciencia Tecnología 2020, 22, 549–558. [Google Scholar] [CrossRef]
  53. Kannan, P.; Biernacki, J.J.; Visco, D.P.; Lambert, W. Kinetics of thermal decomposition of expandable polystyrene in different gaseous environments. J. Anal. Appl. Pyrolysis 2009, 84, 139–144. [Google Scholar] [CrossRef]
  54. Şenocak, A.; Alkan, C.; Karadağ, A. Thermal Decomposition and a Kinetic Study of Poly(Para-Substituted Styrene)s. Am. J. Anal. Chem. 2016, 7, 246–253. [Google Scholar] [CrossRef] [Green Version]
  55. Leclerc, P.; Doucet, J.; Chaouki, J. Development of a microwave thermogravimetric analyzer and its application on polystyrene microwave pyrolysis kinetics. J. Anal. Appl. Pyrolysis 2018, 130, 209–215. [Google Scholar] [CrossRef]
  56. Gringolts, M.L.; Dement’Ev, K.I.; Kadiev, K.M.; Maksimov, A.L.; Finkel’Shtein, E.S. Chemical Conversion of Polymer Wastes into Motor Fuels and Petrochemical Raw Materials (A Review). Pet. Chem. 2020, 60, 751–761. [Google Scholar] [CrossRef]
  57. Poutsma, M.L. Mechanistic analysis and thermochemical kinetic simulation of the pathways for volatile product formation from pyrolysis of polystyrene, especially for the dimer. Polym. Degrad. Stab. 2006, 91, 2979–3009. [Google Scholar] [CrossRef]
  58. Kannan, P.; Biernacki, J.J.; Visco, D.P., Jr. A review of physical and kinetic models of thermal degradation of expanded polystyrene foam and their application to the lost foam casting process. J. Anal. Appl. Pyrolysis 2007, 78, 162–171. [Google Scholar] [CrossRef]
  59. Hussain, Z.; Khan, K.M.; Perveen, S.; Hussain, K.; Voelter, W. The conversion of waste polystyrene into useful hydrocarbons by microwave-metal interaction pyrolysis. Fuel Process. Technol. 2012, 94, 145–150. [Google Scholar] [CrossRef]
  60. Aljabri, N.M.; Lai, Z.; Huang, K.-W. Selective conversion of polystyrene into renewable chemical feedstock under mild conditions. Waste Manag. 2018, 78, 871–879. [Google Scholar] [CrossRef] [PubMed]
  61. Baena-González, J.; Santamaria-Echart, A.; Aguirre, J.L.; González, S. Chemical recycling of plastic waste: Bitumen, solvents, and polystyrene from pyrolysis oil. Waste Manag. 2020, 118, 139–149. [Google Scholar] [CrossRef] [PubMed]
  62. Park, K.-B.; Jeong, Y.-S.; Guzelciftci, B.; Kim, J.-S. Two-stage pyrolysis of polystyrene: Pyrolysis oil as a source of fuels or benzene, toluene, ethylbenzene, and xylenes. Appl. Energy 2020, 259, 114240. [Google Scholar] [CrossRef]
  63. Singh, R.K.; Ruj, B.; Sadhukhan, A.K.; Gupta, P. Thermal degradation of waste plastics under non-sweeping atmosphere: Part 2: Effect of process temperature on product characteristics and their future applications. J. Environ. Manag. 2020, 261, 110112. [Google Scholar] [CrossRef]
  64. Nisar, J.; Ali, G.; Shah, A.; Shah, M.R.; Iqbal, M.; Ashiq, M.N.; Bhatti, H.N. Pyrolysis of Expanded Waste Polystyrene: Influence of Nickel-Doped Copper Oxide on Kinetics, Thermodynamics, and Product Distribution. Energy Fuels 2019, 33, 12666–12678. [Google Scholar] [CrossRef]
  65. Onwudili, J.A.; Insura, N.; Williams, P.T. Composition of products from the pyrolysis of polyethylene and polystyrene in a closed batch reactor: Effects of temperature and residence time. J. Anal. Appl. Pyrolysis 2009, 86, 293–303. [Google Scholar] [CrossRef]
  66. Schröder, U.K.O.; Ebert, K.H.; Hamielec, A.W. On the Kinetics and Mechanism of Thermal Degradation of Polystyrene, 2. Formation of Volatile Compounds. Die Makrom. Chemie 1984, 185, 991–1001. [Google Scholar] [CrossRef]
  67. Carniti, P.; Beltrame, P.L.; Armada, M.; Gervasini, A.; Audisio, G. Polystyrene thermodegradation. 2. Kinetics of formation of volatile products. Ind. Eng. Chem. Res. 1991, 30, 1624–1629. [Google Scholar] [CrossRef]
  68. Zhou, J.; Qiao, Y.; Wang, W.; Leng, E.; Huang, J.; Yu, Y.; Xu, M. Formation of styrene monomer, dimer and trimer in the primary volatiles produced from polystyrene pyrolysis in a wire-mesh reactor. Fuel 2016, 182, 333–339. [Google Scholar] [CrossRef]
  69. Cha, W.S.; Kim, S.B.; McCoy, B.J. Study of polystyrene degradation using continuous distribution kinetics in a bubbling reactor. Korean J. Chem. Eng. 2002, 19, 239–245. [Google Scholar] [CrossRef]
  70. Kruse, T.M.; Woo, O.S.; Broadbelt, L.J. Detailed mechanistic modeling of polymer degradation: Application to polystyrene. Chem. Eng. Sci. 2001, 56, 971–979. [Google Scholar] [CrossRef]
  71. Kruse, T.M.; Woo, O.S.; Wong, H.-W.; Khan, S.S.; Broadbelt, L.J. Mechanistic Modeling of Polymer Degradation: A Comprehensive Study of Polystyrene. Macromolecules 2002, 35, 7830–7844. [Google Scholar] [CrossRef]
  72. Levine, S.E.; Broadbelt, L.J. Reaction pathways to dimer in polystyrene pyrolysis: A mechanistic modeling study. Polym. Degrad. Stab. 2008, 93, 941–951. [Google Scholar] [CrossRef]
  73. Kwak, H.; Shin, H.-Y.; Bae, S.-Y.; Kumazawa, H. Characteristics and kinetics of degradation of polystyrene in supercritical water. J. Appl. Polym. Sci. 2006, 101, 695–700. [Google Scholar] [CrossRef]
  74. Scheirs, J.; Kaminsky, W. Overview of Commercial Pyrolysis Processes for Waste Plastics. In Feedstock Recycling and Pyrolysis of Waste Plastics: Converting Waste Plastics into Diesel and Other Fuels; John Wiley & Sons Ltd.: Chichester, UK, 2006; pp. 383–433. ISBN 978-0-470-02152-1. [Google Scholar]
  75. Veksha, A.; Giannis, A.; Chang, V.W.-C. Conversion of non-condensable pyrolysis gases from plastics into carbon nanomaterials: Effects of feedstock and temperature. J. Anal. Appl. Pyrolysis 2017, 124, 16–24. [Google Scholar] [CrossRef]
  76. Gulati, S.; Harish Neela Lingam, B.; Kumar, S.; Goyal, K.; Arora, A.; Varma, R.S. Improving the air quality with Functionalized Carbon Nanotubes: Sensing and remediation applications in the real world. Chemosphere 2022, 299, 134468. [Google Scholar] [CrossRef]
  77. Miandad, R.; Kumar, R.; Barakat, M.A.; Basheer, C.; Aburiazaiza, A.S.; Nizami, A.S.; Rehan, M. Untapped conversion of plastic waste char into carbon-metal LDOs for the adsorption of Congo red. J. Colloid Interface Sci. 2018, 511, 402–410. [Google Scholar] [CrossRef] [PubMed]
  78. Dogu, O.; Pelucchi, M.; Van de Vijver, R.; Van Steenberge, P.H.; D’Hooge, D.R.; Cuoci, A.; Mehl, M.; Frassoldati, A.; Faravelli, T.; Van Geem, K.M. The chemistry of chemical recycling of solid plastic waste via pyrolysis and gasification: State-of-the-art, challenges, and future directions. Prog. Energy Combust. Sci. 2021, 84, 100901. [Google Scholar] [CrossRef]
Figure 1. EPS pyrolysis experimental scheme.
Figure 1. EPS pyrolysis experimental scheme.
Polymers 14 04957 g001
Figure 2. EPS TGA curve at a heating rate of 20 °C min−1. Measuring points (+).
Figure 2. EPS TGA curve at a heating rate of 20 °C min−1. Measuring points (+).
Polymers 14 04957 g002
Figure 3. Influence of heating rate on EPS performance.
Figure 3. Influence of heating rate on EPS performance.
Polymers 14 04957 g003
Figure 4. Influence of temperature and heating rate on the obtaining of value-added products. Heating rates: (a) 4 °C min−1, (b) 12 °C min−1, (c) 25 °C min−1, (d) 40 °C min−1.
Figure 4. Influence of temperature and heating rate on the obtaining of value-added products. Heating rates: (a) 4 °C min−1, (b) 12 °C min−1, (c) 25 °C min−1, (d) 40 °C min−1.
Polymers 14 04957 g004
Figure 5. Reaction mechanism for the degradation of polystyrene.
Figure 5. Reaction mechanism for the degradation of polystyrene.
Polymers 14 04957 g005
Figure 6. Standardized Pareto diagram for liquid oil yield.
Figure 6. Standardized Pareto diagram for liquid oil yield.
Polymers 14 04957 g006
Figure 7. Main effects on the liquid oil yield.
Figure 7. Main effects on the liquid oil yield.
Polymers 14 04957 g007
Figure 8. Standardized Pareto chart for styrene production.
Figure 8. Standardized Pareto chart for styrene production.
Polymers 14 04957 g008
Figure 9. Main effects on styrene production.
Figure 9. Main effects on styrene production.
Polymers 14 04957 g009
Figure 10. Response surface.
Figure 10. Response surface.
Polymers 14 04957 g010
Table 1. Summary of physicochemical properties of PS raw material.
Table 1. Summary of physicochemical properties of PS raw material.
ParameterValueReferences
Density, kg/m31040–1050[17,32]
Melting point, °C180–260[17,33]
HHV, MJ/kg37.22–42.1[34,35]
Ultimate analysis, wt.%
C66.47–92.7[25,32,34,35,36,37]
H7.4–9.43[25,32,34,35,36,37]
N0–0.8[25,34,35,36,37]
S0–0.51[25,34,35,36,37]
O0–6.8[25,34,35,36,37]
Proximate analysis, wt.%
Moisture0–0.24[25,34,35,36,37]
Volatile88.9–99.59[25,32,34,35,36,37]
Ash0–4.6[25,34,35,36,37]
Fixed carbon0.1–2.25[25,32,34,35,36,37]
Table 2. Influence of temperature on the performance of the EPS pyrolysis process.
Table 2. Influence of temperature on the performance of the EPS pyrolysis process.
Temperature °CYield, wt% 1
Heating Rate: 4 °C min−1Heating Rate: 12 °C min−1
LiquidGasSolidLiquidGasSolid
35076.511.611.98410.645.36
40088.54.756.891.53.694.81
450933.014924.243.76
500923.34.7915.123.88
Temperature °CHeating rate: 25 °C min−1Heating rate: 40 °C min−1
LiquidGasSolidLiquidGasSolid
35082.59.717.882.56.3411.2
40090.54.355.285.58.296.2
450913.98589.55.844.7
500895.665.386.56.746.8
1 Yield variation of ±1%.
Table 3. Physicochemical properties of liquid hydrocarbon from EPS pyrolysis.
Table 3. Physicochemical properties of liquid hydrocarbon from EPS pyrolysis.
EPS Pyrolysis-Derived OilLu et al. [33]Van der Westhuizen et al. [35]
Parameters
Temperature, °C450500420550
Heating rate, °C min−125125n.r.
Yields, wt.%
Liquid hydrocarbon899176.2482.5
Gas5.663.8810.753
Solid5.355.1213.010.4
Styrene72.9971.387339.4
Properties
Density at 15 °C, kg m−3933935n.r.923
Kinematic viscosity, mm2 s−11.091.17n.r.0.88
Heating value, MJ kg−141.6441.65n.r.n.r.
n.r. = not reported.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gonzalez-Aguilar, A.M.; Cabrera-Madera, V.P.; Vera-Rozo, J.R.; Riesco-Ávila, J.M. Effects of Heating Rate and Temperature on the Thermal Pyrolysis of Expanded Polystyrene Post-Industrial Waste. Polymers 2022, 14, 4957. https://doi.org/10.3390/polym14224957

AMA Style

Gonzalez-Aguilar AM, Cabrera-Madera VP, Vera-Rozo JR, Riesco-Ávila JM. Effects of Heating Rate and Temperature on the Thermal Pyrolysis of Expanded Polystyrene Post-Industrial Waste. Polymers. 2022; 14(22):4957. https://doi.org/10.3390/polym14224957

Chicago/Turabian Style

Gonzalez-Aguilar, Arantxa M., Victoria P. Cabrera-Madera, James R. Vera-Rozo, and José M. Riesco-Ávila. 2022. "Effects of Heating Rate and Temperature on the Thermal Pyrolysis of Expanded Polystyrene Post-Industrial Waste" Polymers 14, no. 22: 4957. https://doi.org/10.3390/polym14224957

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