Next Article in Journal
Integrated Steady-State System Package for Nuclear Thermal Propulsion Analysis Using Multi-Dimensional Thermal Hydraulics and Dimensionless Turbopump Treatment
Next Article in Special Issue
Water and Emerging Energy Markets Nexus: Fresh Evidence from Advanced Causality and Correlation Approaches
Previous Article in Journal
Short-Term Load Forecasting for Residential Buildings Based on Multivariate Variational Mode Decomposition and Temporal Fusion Transformer
Previous Article in Special Issue
Processing Orchard Grass into Carbon Bio Pellets via Hydrothermal Carbonisation—A Case Study Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Methods and Validation Techniques of Chemical Kinetics Models in Waste Thermal Conversion Processes

by
Magdalena Skrzyniarz
1,
Marcin Sajdak
2,
Anna Biniek-Poskart
3,
Andrzej Skibiński
3,
Marlena Krakowiak
1,
Andrzej Piotrowski
4,
Patrycja Krasoń
1 and
Monika Zajemska
1,*
1
Faculty of Production Engineering and Materials Technology, Czestochowa University of Technology, 19 Armii Krajowej Ave., 42-200 Czestochowa, Poland
2
Center of New Technologies, Department of Air Protection, Faculty of Energy and Environmental Engineering, Silesian University of Technology, 22 B Konarskiego Ave., 44-100 Gliwice, Poland
3
Faculty of Management, Czestochowa University of Technology, 19 B Armii Krajowej Ave., 42-200 Czestochowa, Poland
4
Faculty of Mechanical Engineering and Computer Science, Czestochowa University of Technology, 42-200 Czestochowa, Poland
*
Author to whom correspondence should be addressed.
Energies 2024, 17(13), 3067; https://doi.org/10.3390/en17133067
Submission received: 19 April 2024 / Revised: 6 June 2024 / Accepted: 19 June 2024 / Published: 21 June 2024
(This article belongs to the Special Issue Bioenergy Economics: Analysis, Modeling and Application)

Abstract

:
This article discusses the potential of using computer-simulation methods in processes such as thermal waste conversion, i.e., pyrolysis, gasification, combustion and torrefaction. These methods are gaining in importance, among others, due to the difficulties in execution and high costs associated with conducting experimental research in real conditions or the need to obtain detailed data on the phenomenon under study in a relatively short time. Computer simulation also allows for numerous errors to be avoided, such as those that may occur during optimization activities, the effects of which may have serious consequences, both economic and environmental. In addition to their many advantages, the limitations and disadvantages of using computer-simulation methods were also indicated, mainly related to the interpretation and validation of the results obtained using modelling. Owing to the complexity of the phenomena occurring during thermal conversion, special attention was focused on models based on chemical kinetics, thanks to which it is possible to predict the quantitative and qualitative composition of products in these processes. The aim of this research is to identify the research gap in the field of issues related to models of chemical kinetics of thermal waste conversion processes.

1. Introduction

Efforts are being made worldwide to promote the energy use of waste through combustion [1], gasification [2], pyrolysis [3] and torrefaction [4] as a strategy that fits, among others, into the idea of a circular economy [5,6]. These activities focus on finding sustainable alternatives for the effective use of waste, which loses the status of “rubbish” and becomes a valuable raw material for other processes [1]. Unfortunately, due to the heterogeneity of waste and the complexity of the phenomena occurring during thermal conversion, there are a number of limitations to its widespread use [7].
The European Union strongly supports activities aimed at preventing waste and reusing products. According to the EU Landfill Directive, by 2035, EU countries must limit the amount of municipal waste sent to landfills to 10% or less of the total municipal waste generated [8]. This creates an opportunity to manage waste that has previously been stored in landfills. Currently, the EU places great emphasis on increasing the share of recycling in waste management methods and reducing the share of landfilling [9].
Examples include two neighboring countries: Poland and Germany. Both countries are striving to transition their economies to a circular economy. In this respect, Germany is considered to be a country that has made a number of far-reaching changes to achieve this goal. Currently, 79% of all waste generated in Germany (64% of municipal waste) is recycled (data from Statistische Bundesamt). In Poland, 36.3% of all waste generated (26.7% of municipal waste) was recycled in 2022—according to data from the Central Statistical Office (called GUS in Poland). Both countries mentioned above have a similar hierarchy of waste management (Figure 1).
Germany has introduced the “Principles for Sustainable Waste Management”, which contributes to the elimination of waste. These rules specify that the producer is responsible for each waste produced (also financially—“the polluter pays”). This encourages manufacturers to create new products responsibly. The “precautionary principle” allows for intervention in areas where general regulations do not function adequately. The “principle of proximity” requires that waste be disposed of as close as possible to the place of its generation, which does not exclude the export of waste if it is economically profitable. The “principle of subsidiarity” determines who will deal with waste management, taking into account factors such as transport costs, efficiency and benefits resulting from the proximity of the place of waste production and processing [11]. A special group of waste is chemical waste that, due to its properties (flammability, explosiveness, toxicity and others), is classified as hazardous waste [12,13,14] and requires proper handling [15], with particular care. Chemical waste includes waste oils [16], waste-containing asbestos, medical and veterinary waste, expired plant protection products and household chemicals [17]. As numerous scientific works indicate, hazardous waste is the most difficult waste to manage because, in the process of its thermal processing, it produces, among others, heavy metals and dioxins. Governments around the world are making a number of efforts to better manage and deal with hazardous waste. An example is Portugal. Over the last decades, the Portuguese government has introduced numerous solutions, such as the modification of the legal framework, creation of a network infrastructure that includes integrated systems of industrial waste treatment (ISIWT) and also development of new organizational methods to make the waste management system more effective [16].
Widely developed computerization addresses these problems by proposing numerical-simulation methods as an important tool in the analysis of complex chemical and gasodynamic phenomena occurring during the thermal conversion of waste [18,19]. The interest in using computer-simulation methods in thermal waste conversion processes is evidenced by numerous scientific publications [20,21]. Using the popular science mapping tool VOSviewer (version 1.6.20) [20,21,22,23], an analysis was carried out using the Scopus database.
The datasets prepared on the basis of the Scopus database include an analysis of global literature on research topics focusing on chemical kinetics in relation to thermal conversion of waste. The research was performed in May 2024. The results of the literature analysis are based on the use of keywords and restrictions introduced in the advanced search in the Scopus database. In order to limit the number of documents found, only the titles of articles were analyzed (not abstracts and keywords). The search with the use of the string “TITLE-ABS-KEY” resulted in finding over a thousand documents that were not necessarily connected with the analyzed subject matter. The used string was as follows: TITLE ((“chemical kinetics” OR “chemical models” OR “kinetics model” OR “kinetic theory” OR “kinetics” OR “kinetic modeling”) AND (“waste”) AND (“thermal conversion” OR “gasification” OR “combustion” OR “torrefaction” OR “pyrolysis”)) AND PUBYEAR > 2018 AND PUBYEAR < 2026 AND (LIMIT-TO (SRCTYPE, “j”)). The number of results found was 212 documents. The timeframe was limited to the last six years, and the searching formula was limited only to article titles that appear in journals. Then, the results selected from the Scopus database were exported as CSV files and supplied into the VOSviewer program in order to analyze the correlations of the obtained literature list for a given subject area. The observed dependencies and relationships between the selected keywords required the creation of several datasets. Visualizations in the form of a map were created using the VOSviewer 1.6.18 software. As part of the attempt to interpret the keywords, the following approach was taken: the type of analysis was “co-occurrence”, the unit of analysis was “All keywords”, and the counting method was “Full counting”. The “minimum number of occurrences of a keyword” was set to six. This led to 142 results as those that “meet the threshold” among 2005 keywords. If the minimal number of occurrences of a keyword was five, the results that meet the threshold were 175. If that number was seven, it led to 107 results, and if the number was eight, the number of results was 89. Two of all the keywords were eliminated due to the lack of their importance to the subject matter, and they were “article” and “priority journal”. Figure 2 as well as Figure 3 show that the most frequently appearing words were kinetics, activation energy and pyrolysis with the total link strength of 1571, 1570 and 1561, respectively. However, the ranking of a keyword’s total link strength may differ a bit from that of their occurrences.
Among a total number of 2005 keywords, there were three that appear with a frequency of over 130 times; one keyword with a frequency of over 80, 60, 50 and 40 times each; and seven keywords with a frequency of over 20 times.
Using the VOSviewer software, the correlation between particular keywords that appear in documents being analyzed can be identified and presented as a map. The words are grouped into clusters to show their relationship. The size of the circles is positively correlated with the occurrence of the keywords. The more often a keyword appears, the larger the font size and circles. The strength of occurrences is indicated by the size of the nodes, while the strength of the relationships reflects the thickness of the lines between the nodes. The analysis shows that five clusters were identified in the network (red, green, blue, purple and yellow). The red cluster is connected with the keyword “kinetics” that indicates 139 links, “activation energy” with 139 links, “thermogravimetric analysis” with 136 links, as well as “biomass” indicating 121 links and “waste incineration” with 102 links. The green cluster refers to the keyword “pyrolysis”, indicating 137 links, correlated with “pyrolysis kinetics”. The connectivity included in the blue, purple and yellow clusters is less dominant, which is presented by the size of their circles.
Analyzing the “network” visualization, it was observed that, over the years, interest in chemical kinetics in the thermal conversion of waste processes has been growing. This is confirmed by the network of connections within the clusters. The keyword maps clearly indicate the high importance of the research topic undertaken, emphasizing the need to expand research in this area, especially in the context of predicting the chemical composition of thermal waste conversion products using chemical kinetics models. The analysis shows that, despite many scientific works published on this topic, there is still a research gap in the issues discussed in the article. With a view to promoting a sustainable development strategy around the world in which the effective use of waste (with diverse properties and elemental composition) plays an important role, there is an urgent need to develop mathematical models that take into account the chemical kinetics of thermal waste conversion. The use of computer-simulation methods in the analysis of complex thermal decomposition processes will assess their energy suitability and the environmental effects of their use.

2. Theoretical Background

2.1. The Essence of Computer Simulations

As a result of the development of technology, computer-simulation methods have been gaining importance in recent years, being used not only in many industries but also in business and government institutions [24]. Simulation takes various forms from modelling simple systems and systems for operational management purposes, such as production and finance, to modelling complex systems for corporations, the global economy or environmental organizations [25,26]. In the initial stages of the development of simulation methods, simulations were used primarily in the natural, technical and economic sciences [23,26,27]. They are currently widely used on a wide scale in various industries, such as the automotive industry [25,28] (Nissan, PSA Peugeot Citroën, Toyota, Ford Motor Company) and fields of science, in particular global scientific and research centers [29] (the Universities of Michigan, Tokyo, Minnesota, Karlsruhe and Colorado) as well as by the National Aeronautics and Space Administration (NASA) [30,31,32]. Currently, there is even a trend of the overuse of computer-simulation methods, which can be explained not only by the degree of development of science but also by the ease and universality of their usage. The appropriate use of computer-simulation methods may be determined by many factors, in particular [23,33] the following:
  • Too high costs of conducting experimental research in real conditions;
  • The need to quickly obtain accurate data on the phenomenon under study;
  • The need to obtain a range of information that is impossible to obtain exponentially, which will complement and enrich the existing state of knowledge;
  • The need to avoid mistakes whose consequences may be serious;
  • The lack of an actual research object or when it is only at the design stage.
Numerical models typically consist of input data, decision variables and output data (Figure 4) [34,35,36,37]. Input data introduce constant values into the model for a single experiment, while decision variables are variables controlled by the researcher. The output data are determined by the input and decision data; they enable the selection of a certain combination of decision variable values that is best from the point of view of the analyzed problem.
The greatest advantage of computer simulations is the ability to analyze various time periods with a full picture of the phenomena occurring at each stage of the analyzed process. This is extremely difficult or even impossible to observe in real conditions. A great advantage of simulation is also the repeatability of processes because, once the model has been developed, it can simulate the process any number of times. In addition, the model can incorporate various variants, such as forcing, interference and other changes. The cost of building a model is, in many cases, much lower than conducting research in real conditions, and the possibilities of interpreting the results increase with the number of recorded parameters. However, it is important to note that both the design and interpretation of simulation results require experience and skills to include all relevant parameters and variables for a given process in the simulated problem. Therefore, the main disadvantage of computer simulations is the possibility of making an error when creating the model and the difficulty in properly interpreting the simulation results. These issues are caused by the human factor since a person creates the simulation model and may incorrectly formulate the simulated problem and adjust the tool as well as draw incorrect conclusions from the simulations (Figure 5) [38,39,40].

2.2. Application of Numerical-Simulation Methods in the Analysis of Thermal Conversion Processes of Fuels and Waste

As numerous scientific works indicate, the main way to produce cleaner energy [41,42], especially in the era of energy transformation and decarbonization [43], is to employ thermal conversion processes such as pyrolysis [44,45,46], combustion [47,48,49], gasification [50,51,52,53,54] or torrefaction [55,56]. The input material in the above-mentioned processes includes not only commonly used biomass [18,57,58] but also various types of waste [59,60], including plastics (e.g., PP, PET, PE) [61,62,63,64], municipal solid waste (MSW) [65], waste from the recycling of scrap tires [66,67], rural domestic waste [68] and refuse-derived fuel (RDF) [69,70,71,72] and solid recovered fuel (SRF) [73]. The initial stage of both combustion and gasification is pyrolysis [74,75].
Numerous scientific and research centers are attempting to increase the share of waste in energy production through thermal conversion [76,77]. However, providing a comprehensive description of thermal degradation is highly problematic due to the complexity of these processes [18,78,79]. The urgent need to understand and optimize these processes [80] has contributed in recent years to intensive development in the elaboration of numerical models [81,82], and the main research efforts are focused on understanding the reaction mechanisms of the kinetics of thermal conversion processes [83,84,85]. Chemical kinetics is crucial in comprehending the complex chemistry of the reactions of these processes and the creation of mathematical models [86,87,88,89].
Thermal conversion processes can be simulated using three different approaches:
i.
Computational fluid dynamics models [47,90,91,92,93];
ii.
Kinetic models [18,94,95];
iii.
Equilibrium models [52,96,97,98].
Equilibrium models are generally easy to implement, among others, due to numerous simplifications [99] (Figure 6). Depending on the method used to calculate chemical equilibrium, equilibrium models can be classified into two categories:
i.
Stoichiometric models;
ii.
Non-stoichiometric models.
In stoichiometric models, the equilibrium is determined by using equilibrium constants for each reaction involved in the analyzed process. In contrast, in non-stoichiometric models, it is determined by minimizing the Gibbs free energy [38].
Kinetic modelling is quite complex [100], and the models are divided into the following:
Most of the models described in the literature are simplified models that only allow for the prediction of the char yield and degassing time [108,109]. Unfortunately, this approach is not sufficient. From the energetic point of view of thermal conversion processes, it is extremely important to provide information on the temperature distribution, kinetic parameters and chemical composition of the gases released during the above-mentioned processes [98,110]. Although these issues have been extensively studied both theoretically and experimentally using various techniques, they still require more attention and research [111,112,113]. Thermal analysis techniques, such as (TG) [107], (DSC), (DTA), FTIR [114], MS or various combinations thereof, i.e., (TG–FTIR) [115], (TG–MS) or (TG–MS–FTIR), are considered effective techniques for examining the thermal degradation of both fuels and waste. The above-mentioned techniques provide information about the analyzed processes, enabling a better understanding of them, and above all, they can constitute input data for computer simulations [116,117,118].
Table 1 lists the thermal conversion processes in which computer simulations were employed.
According to the analysis, researchers have most often used the equation describing the relationship to explain the phenomena occurring during the thermal conversion of biomass and, in particular, pyrolysis (1) [104]:
B i o m a s s   ( S B ) K ( t ) V o l a i t l e c o n d e n s a b l e + n o n   c o n d e n s a b l e + B i o c h a r
Che et al. [137] noted that Huazhong University of Science and Technology State Key Laboratory of Coal Combustion made significant advancements in the numerical simulation of gasification technology. Yang Haiping utilized the HSC chemistry package and the PSR reactor model to simulate the thermodynamic equilibrium and dynamic equilibrium of the palm oil waste pyrolysis process. The results obtained by Haiping revealed that the initial products contain significant amounts of H2, CO, CO2, CH4, H2O and coke and small amounts of C2H2, C2H6, C3H8 and others in addition to the concentrations of H2 and CO, which grow with the rise in the pyrolysis temperature.
Noteworthy are the computer simulations conducted by Peters et al. [138], who modelled the biomass pyrolysis process using an innovative reaction model in the Aspen Plus (version v14) program. This model was based on the three main building blocks of woody biomass, namely cellulose, hemicellulose and lignin. The pyrolysis process was described using 149 reactions. The linear regression algorithm took into account secondary pyrolysis reactions, enabling the calculation of slow and intermediate pyrolysis reactions. The model of bio-oil included chemical compounds, such as organic acids, aldehydes, alcohols, ketones, phenols, sugar derivatives and degraded lignin, and the produced carbon was modelled implementing the actual elemental composition.
A diagram of the reaction mechanism employed by Petersen et al. is presented in Figure 7 [138].
The above model takes into account primary pyrolysis reactions as well as secondary cracking reactions. The pyrolysis mechanism was divided into three phases, namely one decomposition phase and two pyrolysis phases. The first phase is a virtual reaction step that breaks down the biomass into three main biochemical building blocks: cellulose, hemicellulose and lignin. During the second phase, the decomposition and volatilization of biomass fragments took place, leading to high efficiency of the liquid fraction. This is the dominant reaction mechanism for fast pyrolysis processes with short gas residence times. The third phase involved secondary cracking and charring reactions, which increased the gas and carbon yields at the expense of liquid yields as a consequence of secondary (catalytic) cracking reactions. These reactions are more significant with the increasing residence time and are, therefore, particularly important for slow and intermediate pyrolysis reactions. Petersen et al. [138] proposed a model that allows for the simulation of fast and slow pyrolysis of any lignocellulosic raw material with a known composition [138]. Moreover, a predictive and detailed modelling approach enabled a comprehensive assessment of the properties of bio-oils obtained from various types of lignocellulosic biomass under various pyrolysis conditions (including fast and slow pyrolysis). Peters et al. emphasize that, to date, the analysis of pyrolysis processes has been based on simple models with a very simplified composition of bio-oil. The authors point out that their work is the first to present a comprehensive kinetic model of the reaction, which can be easily implemented in Aspen Plus and similar thermal conversion simulation software packages.
A similar approach was proposed by Granados et al., who modelled the biomass torrefaction process consisting of the three stages previously described in Petersen’s work [139]. The process starts with drying the biomass to remove moisture, followed by heating the completely dry biomass. When the temperature grows, the biomass degassing process begins, resulting in the formation of condensable and non-condensable volatile substances and a solid residue. At low temperatures, the hemicellulose decomposition rate is high and dominates the entire torrefaction process. Only at high temperatures do some of the cellulose and lignin also begin to decompose but to a lesser extent. After primary degassing, secondary reactions occur where the produced condensable volatiles can react and form new solid and volatile components [139].
Park et al. [140] proposed a simplified approach to modelling the thermal conversion of biomass. Using Aspen Plus (version 14.3) software, they modelled the impact of the process operating conditions on biomass torrefaction using a one-dimensional reactor. The input data for the modelling were taken from the literature, and the diagram of the kinetic mechanism is presented in Figure 8. The results obtained by Park et al. confirmed that the energy efficiency of the torrefaction process increases with temperature (Figure 8).
Yet another approach was proposed by Bates et al. [141], who developed a kinetic model for estimating the torrefaction products. The model is based on the mechanism originally proposed by Di Blasi and Lanzetta for the degradation of pure hemicellulose. Oliviera et al. [142] developed a kinetic model of the solid waste gasification process. The proposed mathematical model allowed for the detailed analysis of the processes occurring during gasification, taking into account partial derivatives, through the use of a CFD tool.
Also noteworthy is the approach to modelling the thermal conversion of biomass proposed by Wang et al. [126]. The authors developed a model for predicting pyrolysis products based on the Coats–Redfern method and the Kissinger method (Figure 9).
Yet another approach to simulating the biomass gasification process was proposed in the work by Gambarotta [42], who used a non-stoichiometric equilibrium model in downdraft gasifiers. The author utilized a five-component composition of the raw material for the calculations, i.e., forest waste (carbon, hydrogen, oxygen, nitrogen and sulfur), and included fifteen chemical compounds for the synthesis gas. The model calculated the relative amounts of gasification products and the calorific value of the synthesis gas. Gambarotta emphasized that the advantage of the non-stoichiometric approach is that it can easily calculate not only the concentrations of the main gasification products but also the concentrations of the by-products, especially chemical compounds containing nitrogen and sulfur. At the same time, he pointed out one of the main disadvantages of equilibrium models, namely the overestimation of molecular hydrogen and the underestimation of methane. The model developed by Gambarotta was analyzed for the sensitivity of the influence of the selected parameters on the process and then verified by comparing the obtained simulation results with experimental data found in the literature. The model occurred to be effective in simulating the biomass gasification process with an acceptable level of error.
The actual reaction patterns of waste pyrolysis, including biomass and plastics, are extremely complex due to the formation of over one hundred intermediate products [117,143,144]. RDF is a special type of waste that is more complex than biomass due to its heterogeneous nature [72]. The amount of research on the kinetic modelling of RDF is much smaller compared to other fuels [71,145,146]. Three different approaches have been used in the literature to model the kinetics of thermal degradation of complex solid fuels, such as RDF. The first one considers fuel as one homogeneous species. The second approach considers the RDF pyrolysis rate as a weighted sum of the pyrolysis rates of cellulose, lignin, hemicellulose, polyethylene and other plastics. The third approach, similar to the second, considers the RDF pyrolysis rate as the weighted sum of the pyrolysis rates of paper, cardboard, wood, polyethylene and other plastics. The second and third approaches assume that possible interactions between the RDF components have a negligible impact on pyrolysis [71,146].
The research conducted by Zhou et al. [144] deserves attention. The authors developed a model to describe the heat transfer processes inside a porous RDF particle composed of polyethylene and cardboard. The model also considers the heat of the fusion of polyethylene (PE). To describe the degradation of polyethylene and cardboard, the authors adopted a one-stage global model and a scheme of two consecutive first-order reactions. To match the results of experimental research, the authors selected and modified two widely used heat transfer models—the Kunii and Smith model and the Breitbach and Barthels model. Moreover, they investigated the influence of the heat transfer model on the prediction of the RDF pyrolysis process. The authors achieved a high level OF agreement between the results obtained through modelling and the experimental results for both the modified Kunia and Smith model and for the Breitbach and Barthels model.
On the other hand, Çepelioğullar et al. [146] modelled the pyrolysis process of RDF fuel in four reaction stages. The authors observed that the thermal decomposition of RDF involved many stages, which consisted of the decomposition of cellulose and plastics. Similarly, Bosmans et al. [147], using MATLAB (version R2023a) software, modelled the RDF fuel pyrolysis process, assuming four independent, parallel first-order reactions. The model adopted for the calculations enabled high compliance and a good fit with the experimental data since it was able to capture both the degradation of the lignocellulosic fraction and the plastic fraction in the waste material. The authors recommend that researchers conduct TGA experiments on RDF samples to obtain a full picture of the material degradation process. If RDF was produced from household waste, the DTG degradation profile of this waste may match those reported in the literature. Nonetheless, if RDF was created from different waste streams and was heterogeneous, it is unlikely that a similar DTG profile would be found in the literature. By combining the obtained pyrolysis kinetics with heat and mass transfer, accurate simulation models for real thermochemical conversion systems can be obtained [147].

2.3. Tools for Numerical Simulation of Thermal Conversion Processes of Fuels and Waste

Among the most frequently mentioned programs for simulating thermal conversion processes, the literature on the subject mentions two commercial software programs: Aspen Plus and Fluent (version 2021 R2). It is emphasized that better simulation results can be obtained by using these programs simultaneously [1,148]. Few works, including those of the authors of the present article, mention Ansys Chemkin-Pro (version 2021R1) software as an irreplaceable tool for predicting the gaseous products of thermal conversion (Figure 10) [147,149].
To avoid the complexity of the thermal conversion process, e.g., gasification, and to facilitate the creation of models, some researchers have developed models based on the Aspen Plus process simulator. This involves assessing and analyzing each stage of the process, such as gasification, before integrating it into the global process [149,150,151,152]. Each stage of the process, e.g., gasification, can be described and tested in a block before it is integrated into the global process [153,154]. Simulation calculations using this tool are performed using the built-in database of the simulator’s physical properties and the database of the properties of the chemical components used in the calculations (e.g., biomass). In addition, custom Fortran or Excel subroutines can be included if necessary [81]. The program uses a sequential modular approach to solve the process diagram module-by-module and calculate the outlet stream properties based on the inlet stream properties for each block. In modelling, for example, a gasifier in Aspen Plus, the entire process is divided into several sub-processes, including drying and pyrolysis, oxidation and gasification or reduction.
In recent years, there has been a significant increase in the use of Aspen Plus models for thermal conversion processes, particularly gasification, compared to equilibrium models [150,151,152]. The literature review shows that most researchers rely on non-stoichiometric methods when modelling equilibrium using the Aspen Plus program. On the other hand, stoichiometric equilibrium models are rarely used in the literature, where only numerical models with fewer reactions are described. González-Vázquez et al. [81] compared a non-stoichiometric model and a stoichiometric model, both built using Aspen Plus, to determine their suitability for predicting steady-state biomass gasification efficiency. The simulation results were confirmed by experimental data previously obtained in a semi-pilot atmospheric pressure fluidized bed gasification unit. The authors used Aspen Plus V.8.6 to model biomass gasification and predict synthesis gas. The diagram presents each stage of the process, showing the flows of materials and energy are determined by means of operating blocks.
Models developed with the Ansys Chemkin-Pro tool, which enables accurate modelling of the thermal conversion reaction, e.g., biomass gasification, are definitely popular. The software offers users the ability to develop detailed reaction mechanisms and simulate the thermal decomposition of various raw materials, e.g., biomass, for various process conditions. Additionally, the tool can be integrated with computational fluid dynamics (CFD) models. According to the researchers, Chemkin-Pro is definitely more difficult to use than Aspen Plus, primarily as a consequence of the chemical approach to the analyzed process, based on chemical kinetics, which requires a deeper understanding of the chemical reactions and mechanisms, especially when describing complex biomass gasification processes. Additionally, the technical nature of the software makes it less user-friendly than Aspen Plus, which is designed for a wider range of chemical processes and has a more user-friendly interface. These factors make Chemkin-Pro require a higher level of expertise and effort to effectively utilize its chemical kinetic modelling capabilities in biomass gasification [99].
Examples of the use of Chemkin-Pro software are the simulations carried out by Sieradzka et al. [140] and Slefarski et al. [150]. The authors modelled the chemical composition of the gaseous products of the pyrolysis of fuels derived from waste and identified reaction paths based on kinetic models.

2.4. Validation Techniques of Chemical Kinetics Models in Waste Thermal Conversion Processes

To validate the results of the simulations of the thermal conversion processes of fuels and waste, experimental analyses are primarily used [151,152], including thermogravimetric analysis [111,153,154,155] with a Fourier transform infrared spectrometer (TG–FTIR) [156,157,158] and gas chromatography techniques [75,159,160,161]. Researchers may also compare their results with literature data [52,138].
One of the methods utilized to validate the results obtained by modelling is also statistical analysis employing statistical computing and software [162,163,164]: MATLAB, ANOVA (tool in OriginPro 2024b) version [165,166,167,168] and MS Excel (version 2021).
Statistical approaches [163,164,165,169], such as response surface methodology (RSM) and Taguchi, are commonly used to optimize experimental parameters (Figure 11). In the RSM optimization process, more series of experiments are usually performed than in the Taguchi method. The RSM model outperforms Taguchi in terms of precision and reliability, while Taguchi’s approach is ideal for low-cost and rapid experimental studies. Researchers can choose between the RSM model and Taguchi’s approach based on the complexity and financial resources required for their experiments. Several approaches to RSM have been developed over the years. Box Behnken design (BBD) and central composite design (CCD) technologies are the most commonly utilized RSM technologies for catalytic pyrolysis applications. Professor Taguchi developed optimization techniques (Taguchi techniques), which are commonly used to determine the optimal combination of parameters and their impact on the objective function. Taguchi’s optimization approach often uses the variance test, chi-square test and Fisher’s exact test [170].
Regarding the Taguchi method and RSM, the literature has shown that both methods can provide satisfactory optimization results for selected applications, e.g., thermoelectric generators. However, RSM can provide a more comprehensive optimization analysis because it can show the influence of the parameters on the objective function and the interaction between the parameters. Moreover, in RSM, it is possible to visualize the results of the experimental analysis. The Taguchi method is unique in its ability to provide accurate optimization results with a small number of experiments, making it more cost-effective. The comparison and the most important steps of statistical optimization using the Taguchi, ANNOVA and RSM methods are presented in Figure 12 and Figure 13 [165].

3. Results

As mentioned earlier, experimental analysis is one of the most frequently used methods to verify the obtained simulation results.
Peters et al. [138] compared the obtained simulation results, described in Section 2.2, with literature data, showing good agreement (Figure 14). Additionally, the authors conducted their pyrolysis experiments to validate the reaction model. The raw material used for the research came from beech wood and was subjected to pyrolysis under precisely defined conditions and at various temperatures. The result of the research was the determination of process efficiency in terms of the received products and their chemical composition. In the final stage, the composition of bio-oil obtained by simulating the pyrolysis of beech wood (moisture 8.9%) was compared with literature data, which showed that beech wood had a lower water content. Comparable proportions were found for organic acids and ketones, while much higher disproportions were obtained for alcohol and aldehyde. In the literature cited by the authors, as much as 44% of the bio-oil was not identified [138].
Ali et al. [171] conducted similar calculations using Aspen Plus software to model the chemical composition of the solid and liquid gasification products of a biomass and coal mixture, receiving results similar to those obtained in experimental studies (Figure 15, Figure 16 and Figure 17).

4. Discussion

In the era of energy transformation and decarbonization, many efforts are being made worldwide to promote the use of various types of waste as energy sources. These activities are aimed not only at searching for but also promoting sustainable alternatives, thanks to which waste will become a valuable raw material for other processes. Unfortunately, the heterogeneity and varied chemical composition of waste as well as the complexity of the phenomena occurring during thermal conversion make a comprehensive description of the degradation process difficult. This translates into limitations in their widespread use. Computer-simulation methods have been developed in recent years to address these issues, contributing to the intensification of research in the development of numerical models. The main efforts of many research centers around the world are primarily focused on understanding the gas dynamics and chemical mechanisms occurring during the thermal conversion of waste and, in particular, the reaction kinetics. The literature analysis carried out in this article shows that two commercial software products, Aspen Plus and Fluent, are most often used to simulate the processes of the thermal conversion of waste, while only a few works indicate Ansys Chemkin-Pro software as an irreplaceable tool for predicting the gaseous products of thermal conversion [147,149].
To summarize the topic discussed in this article, the following can be noted:
  • The analysis carried out with the use of the VOSviewer software shows that, despite many scientific works published on this topic, there is still a research gap in the issues discussed in the article.
  • The literature on the subject emphasizes that a number of factors are responsible for the proper use of computer simulation methods, and the most important advantage is the ability to simulate phenomena that are difficult or impossible to implement in real conditions.
  • The results of numerical modeling using simplified chemical mechanisms (CFD modeling) describing the combustion process are subject to large errors (due to the long calculation time, simplifications of chemical kinetics are used in the computational procedure, namely reducing the chemistry of methane oxidation to a single-stage mechanism in which combustion products are CO2, H2O, O2 and N2), which is why they often do not coincide with actual measurements. The assumption of a simple combustion mechanism in the calculations results in significant discrepancies with the measurement results.
  • Due to the possibility to create a multi-stage, complex model in the ANSYS Chemkin-Pro program, which does not use simplifications in thermodynamics and chemical kinetics, various waste thermal processing processes can be modeled with high accuracy.
  • The cost of building a model is often much lower than conducting experimental research, resulting in significant savings in time and money.
  • Computer simulations also have several disadvantages, the most important of which is the possibility of making an error when creating a model that omits parameters and chemical reactions that are important from the point of view of the analyzed process.
  • Moreover, numerous works indicate difficulties and errors when interpreting the obtained simulation results. The above-mentioned defects are the responsibility of humans, who may not only incorrectly formulate the simulated problem and adjust the tool but may also draw incorrect conclusions from the performed simulations.
  • It is crucial to validate the developed model and verify the obtained calculations through experimental means using statistical analysis (e.g., MATLAB, ANOVA) or based on literature data. Without such validation, the results of the model calculations remain hypothetical.
  • The issue of managing environmentally burdensome gaseous products of thermal conversion of calorific waste is of an applied nature, as evidenced by the interest of both entrepreneurs from the waste and metallurgical industries. However, this direction requires extensive theoretical studies in the analysis of chemical mechanisms.
  • Knowledge of chemical mechanisms describing the process of combustion or co-combustion of gases from the thermal conversion of waste will create the possibility of their energy utilization.
  • Providing comprehensive knowledge about combustion kinetics will also enable taking appropriate actions to manage gases after thermal conversion of waste while maintaining the proper operation of the heating chamber and minimizing pollution.
  • Currently, it is particularly important and problematic to properly identify the chemical composition of exhaust gases while, at the same time, focusing on activities aimed at minimizing harmful pollutants, such as PAHs.
The analysis conducted in this article indicates enormous progress in modelling the processes of the thermal conversion of waste This suggests that numerical simulations will likely replace costly experimental research in the future. Nonetheless, it is important to consider that, as the number of phenomena to be modelled increases, the accuracy of the model also increases, resulting in a better reflection of reality, but, at the same time, requiring a longer calculation time and greater computing power, leading to higher costs. Moreover, thermal management of waste towards the recovery of valuable products, such as pyrolysis gas, biochar and condensate, as is the case in the pyrolysis process, is an additional opportunity to improve energy efficiency. Using the potential of post-process gases, especially gases from the pyrolysis process, e.g., of biomass, will have an economic impact, contributing to lower fuel consumption.

5. Conclusions

To sum up, it is important to note that, despite the disadvantages and difficulties associated with creating advanced simulation projects, they are an integral element of the development of all industries. The use of simulation in thermal waste conversion processes is an opportunity to employ them on a wider scale, thanks to a better understanding of the phenomena occurring during their degradation, both gasodynamic and kinetic. The analysis performed using the VOSviewer software indicates the need to develop mathematical models used when modeling thermal processes. Reducing the expenditure that would be necessary for extensive experimental research will contribute to a faster transformation of a linear economy into a circular economy that promotes the efficient use of waste.

Author Contributions

Conceptualization, M.S. (Magdalena Skrzyniarz), M.S. (Marcin Sajdak) and M.Z.; writing—original draft preparation, M.S. (Magdalena Skrzyniarz), M.Z., M.S. (Marcin Sajdak), M.K., A.P., A.B.-P. and A.S.; writing—review and editing, M.S. (Magdalena Skrzyniarz), M.Z., M.S. (Marcin Sajdak), M.K., A.B.-P., A.S., A.P. and P.K.; supervision, M.Z.; project administration, M.S. (Magdalena Skrzyniarz) and M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kuo, W.C.; Lasek, J.; Słowik, K.; Głód, K.; Jagustyn, B.; Li, Y.H.; Cygan, A. Low-Temperature Pre-Treatment of Municipal Solid Waste for Efficient Application in Combustion Systems. Energy Convers. Manag. 2019, 196, 525–535. [Google Scholar] [CrossRef]
  2. Lombardi, L.; Carnevale, E.; Corti, A. Analysis of Energy Recovery Potential Using Innovative Technologies of Waste Gasification. Waste Manag. 2012, 32, 640–652. [Google Scholar] [CrossRef] [PubMed]
  3. Shen, Y.; Chen, X.; Ge, X.; Chen, M. Thermochemical Treatment of Non-Metallic Residues from Waste Printed Circuit Board: Pyrolysis vs. Combustion. J. Clean. Prod. 2018, 176, 1045–1053. [Google Scholar] [CrossRef]
  4. Albergaria Campos, A.M.; Khozhanov, N.; Assis, P.S.; Tursunbaev, K.; Masatbayev, M. Economic and Environmental Analyses of Biomass Torrefaction for Injection as Pulverized Material in Blast Furnaces. REM-Int. Eng. J. 2021, 74, 471–482. [Google Scholar] [CrossRef]
  5. Pan, S.Y.; Du, M.A.; Huang, I.T.; Liu, I.H.; Chang, E.E.; Chiang, P.C. Strategies on Implementation of Waste-to-Energy (WTE) Supply Chain for Circular Economy System: A Review. J. Clean. Prod. 2015, 108, 409–421. [Google Scholar] [CrossRef]
  6. Rathore, P.; Sarmah, S.P. Economic, Environmental and Social Optimization of Solid Waste Management in the Context of Circular Economy. Comput. Ind. Eng. 2020, 145, 106510. [Google Scholar] [CrossRef]
  7. Lu, P.; Huang, Q.; Bourtsalas, A.C.; Themelis, N.J.; Chi, Y.; Yan, J. Review on Fate of Chlorine during Thermal Processing of Solid Wastes. J. Environ. Sci. 2019, 78, 13–28. [Google Scholar] [CrossRef]
  8. Available online: https://www.Europarl.Europa.Eu/Topics/Pl/Article/20180328STO00751/Zrownowazone-Zarzadzanie-Odpadami-Dzialania-Ue (accessed on 22 May 2024).
  9. European Parliament; Council of the European Union. Dyrektywa Parlamentu Europejskiego i Rady 2008/98/WE Dnia 19 Listopada 2008 r.W Sprawie Odpadów Oraz Uchylająca Niektóre Dyrektywy. Dziennik Urzędowy Unii Europejskiej 2008, 22, 3–30. [Google Scholar]
  10. Available online: https://www.Bonnorange.de/Nachhaltigkeit/Klimarechner/Abfallhierarchie (accessed on 23 May 2024).
  11. Available online: https://www.Retech-Germany.Net/Themen/Der-Weg-Zur-Modernen-Abfallwirtschaft/Prinzipien-Nachhaltiger-Abfallwirtschaft (accessed on 23 May 2024).
  12. Li, Y.; Fu, Z.; Li, J. Assessing the Policy Benefits of Constructing “Zero-Waste Cities” in China: From the Perspective of Hazardous Waste Lifecycle Management. Sci. Total. Environ. 2024, 918, 170184. [Google Scholar] [CrossRef]
  13. Zhang, Z.; Malik, M.Z.; Khan, A.; Ali, N.; Malik, S.; Bilal, M. Environmental Impacts of Hazardous Waste, and Management Strategies to Reconcile Circular Economy and Eco-Sustainability. Sci. Total. Environ. 2022, 807, 150856. [Google Scholar] [CrossRef]
  14. Xi, B.; Yang, T.; Zhao, R.; Jing, L.; Gong, T.; Huang, Q.; Hou, L. Hazardous Waste Management in the Guangdong–Hong Kong–Macao Greater Bay Area. Engineering 2022, 8, 25–28. [Google Scholar] [CrossRef]
  15. Andrade, L.C.; Míguez, C.G.; Gómez, M.C.T.; Bugallo, P.M.B. Management Strategy for Hazardous Waste from Atomised SME: Application to the Printing Industry. J. Clean. Prod. 2012, 35, 214–229. [Google Scholar] [CrossRef]
  16. Couto, N.; Silva, V.; Monteiro, E.; Rouboa, A. Hazardous Waste Management in Portugal: An Overview. Energy Procedia 2013, 36, 607–611. [Google Scholar] [CrossRef]
  17. Inglezakis, V.J.; Moustakas, K. Household Hazardous Waste Management: A Review. J. Environ. Manag. 2015, 150, 310–321. [Google Scholar] [CrossRef] [PubMed]
  18. Ranzi, E.; Corbetta, M.; Manenti, F.; Pierucci, S. Kinetic Modeling of the Thermal Degradation and Combustion of Biomass. Chem. Eng. Sci. 2014, 110, 2–12. [Google Scholar] [CrossRef]
  19. Chavando, J.A.M.; Silva, V.; Puig-gamero, M.; Cardoso, J.S.; Tarelho, L.A.C.; Eusébio, D. Simulation of Biomass to Syngas: Pyrolysis and Gasification Processes; Elsevier: Amsterdam, The Netherlands, 2022; ISBN 9780323918794. [Google Scholar]
  20. Li, J.; Suvarna, M.; Li, L.; Pan, L.; Pérez-Ramírez, J.; Ok, Y.S.; Wang, X. A Review of Computational Modeling Techniques for Wet Waste Valorization: Research Trends and Future Perspectives. J. Clean. Prod. 2022, 367, 133025. [Google Scholar] [CrossRef]
  21. Escalante, J.; Chen, W.H.; Tabatabaei, M.; Hoang, A.T.; Kwon, E.E.; Andrew Lin, K.Y.; Saravanakumar, A. Pyrolysis of Lignocellulosic, Algal, Plastic, and Other Biomass Wastes for Biofuel Production and Circular Bioeconomy: A Review of Thermogravimetric Analysis (TGA) Approach. Renew. Sustain. Energy Rev. 2022, 169, 112914. [Google Scholar] [CrossRef]
  22. Khan, S.A.; Ali, I.; Naqvi, S.R.; Li, K.; Mehran, M.T.; Khoja, A.H.; Alarabi, A.A.; Atabani, A.E. Investigation of Slow Pyrolysis Mechanism and Kinetic Modeling of Scenedesmus Quadricauda Biomass. J. Anal. Appl. Pyrolysis 2021, 158, 105149. [Google Scholar] [CrossRef]
  23. Chen, Z.; Hammad, A.W.A. Mathematical Modelling and Simulation in Construction Supply Chain Management. Autom. Constr. 2023, 156, 105147. [Google Scholar] [CrossRef]
  24. Zajemska, M.; Poskart, A. Mozliwości Zastosowania Metod Numerycznych Do Przewidywania I Ograniczania Emisji Zanieczyszczeń z Instalacji Spalania Stosowanych w Przemyśle Chemicznym I Rafineryjnym. Przem. Chem. 2013, 92, 357–361. [Google Scholar]
  25. Kurebwa, J.; Mushiri, T. Design and Simulation of an Integrated Steering System for All-Purpose Sport Utility Vehicles (SUVs)—Case for Toyota. Procedia Manuf. 2019, 35, 56–74. [Google Scholar] [CrossRef]
  26. Hamlaci Baskaya, Y.; Kurt, G.; İlcioğlu, K.; Turan, Z. Assessment of Efficacy of Two Different Simulation Techniques Used in Breech Birth Management Training: A Randomized Controlled Study. Clin. Simul. Nurs. 2024, 87, 101499. [Google Scholar] [CrossRef]
  27. Liu, Z.; Wu, S.; Zuo, H.; Lin, J.; Zheng, H.; Lei, H.; Yu, Q.; Wu, X.; Guo, Z. Freeze-drying pretreatment of watermelon peel to improve the efficiency of pectin extraction: RSM optimization, extraction mechanism, and characterization. Int. J. Biol. Macromol. 2023, 249, 125944. [Google Scholar] [CrossRef] [PubMed]
  28. Louback, E.; Biswas, A.; Machado, F.; Emadi, A. Review Article A Review of the Design Process of Energy Management Systems for Dual-Motor Battery Electric Vehicles. Renew. Sustain. Energy Rev. 2024, 193, 114293. [Google Scholar] [CrossRef]
  29. Baker, K.R.; Liljegren, J.; Valin, L.; Judd, L.; Szykman, J.; Millet, D.B.; Czarnetzki, A.; Whitehill, A.; Murphy, B.; Stanier, C. Photochemical Model Representation of Ozone and Precursors during the 2017 Lake Michigan Ozone Study (LMOS). Atmos. Environ. 2023, 293, 119465. [Google Scholar] [CrossRef]
  30. Tao, Z.; Kawa, S.R.; Jacob, J.P.; Liu, D.Y.; Collatz, G.J.; Wang, J.S.; Ott, L.E.; Chin, M. Application of NASA-Unified WRF Model to Carbon Dioxide Simulation- Model Development and Evaluation. Environ. Model. Softw. 2020, 132, 104785. [Google Scholar] [CrossRef]
  31. Kahre, M.A.; Haberle, R.M.; Wilson, R.J.; Urata, R.A.; Steakley, K.E.; Brecht, A.S.; Bertrand, T.; Kling, A.; Batterson, C.M.; Hartwick, V.; et al. The NASA Ames Legacy Mars Global Climate Model: Radiation Code Error Correction and New Baseline Water Cycle Simulation. Icarus 2023, 400, 115561. [Google Scholar] [CrossRef]
  32. Huff, J.L.; Poignant, F.; Rahmanian, S.; Khan, N.; Blakely, E.A.; Britten, R.A.; Chang, P.; Fornace, A.J.; Hada, M.; Kronenberg, A.; et al. Galactic Cosmic Ray Simulation at the NASA Space Radiation Laboratory—Progress, Challenges and Recommendations on Mixed-Field Effects. Life Sci. Space Res. 2023, 36, 90–104. [Google Scholar] [CrossRef] [PubMed]
  33. Marschik, C.; Roland, W.; Löw-Baselli, B.; Steinbichler, G. Application of Hybrid Modeling in Polymer Processing. Annu. Tech. Conf.-ANTEC Conf. Proc. 2020, 2, 535–542. [Google Scholar]
  34. Zhang, Y.; Ji, Y.; Qian, H. Progress in Thermodynamic Simulation and System Optimization of Pyrolysis and Gasification of Biomass. Green Chem. Eng. 2021, 2, 266–283. [Google Scholar] [CrossRef]
  35. Vikram, S.; Rosha, P.; Kumar, S. Recent Modeling Approaches to Biomass Pyrolysis: A Review. Energy Fuels 2021, 35, 7406–7433. [Google Scholar] [CrossRef]
  36. Irfan, M.; Nabi, R.A.U.; Hussain, H.; Naz, M.Y.; Shukrullah, S.; Khawaja, H.A.; Rahman, S.; Ghanim, A.A.J.; Kruszelnicka, I.; Ginter-Kramarczyk, D.; et al. Response Surface Methodology Analysis of Pyrolysis Reaction Rate Constants for Predicting Efficient Conversion of Bulk Plastic Waste into Oil and Gaseous Fuels. Energies 2022, 15, 9594. [Google Scholar] [CrossRef]
  37. Ramanathan, A.; Begum, K.M.M.S.; Pereira, A.O.; Cohen, C. Biomass Pyrolysis System Based on Life Cycle Assessment and Aspen plus Analysis and Kinetic Modeling. In A Thermo-Economic Approach to Energy from Waste; Elsevier: Amsterdam, The Netherlands, 2022; ISBN 9780128243572. [Google Scholar]
  38. Agh, O. Metody Symulacji Komputerowej. Available online: http://www.pi.zarz.agh.edu.pl/inne/odlewn/Wyklady/PSIZ%20Symulacja.pdf (accessed on 5 January 2024).
  39. Jąderko, K.; Białecka, B. Model Technologiczno-Logistyczny Logistyczny Procesu Energetycznego Procesu Energetycznego Wykorzystania Odpadów; 2016; ISBN 9788365265081. Available online: http://www.stegroup.pl/attachments/article/1/Monografia%20J.pdf (accessed on 6 January 2024).
  40. Zajemska, M. Wymagania Stawiane Technice Obliczeniowej W Zakresie Numerycznego Modelowania Składu Chemicznego Produktów Spalania. Model. Inżynierskie 2011, 41, 453–461. [Google Scholar]
  41. Jaiswal, S.; Sahani, S.; Shar, Y.C. Enviro-Benign Synthesis of Glycerol Carbonate Utilizing Bio-Waste Glycerol over Na-Ti Based Heterogeneous Catalyst: Kinetics and E- Metrics Studies. J. Environ. Chem. Eng. 2022, 10, 107485. [Google Scholar] [CrossRef]
  42. Xiao, H.; Li, Z.; Jia, X.; Ren, J. Waste to Energy in a Circular Economy Approach for Better Sustainability: A Comprehensive Review and SWOT Analysis. In Waste-to-Energy; Academic Press: Cambridge, MA, USA, 2020; ISBN 9780128163948. [Google Scholar]
  43. Zajemska, M.; Sajdak, M.; Iwaszko, J.; Skrzyniarz, M.; Biniek-Poskart, A.; Skibiński, A.; Maroszek, A. The Role of Calorific Waste in Transformation of Iron and Steel Industry towards Sustainable Production. Resour. Conserv. Recycl. 2023, 191, 2022–2024. [Google Scholar] [CrossRef]
  44. Yahya, S.A.; Iqbal, T.; Omar, M.M.; Ahmad, M. Techno-Economic Analysis of Fast Pyrolysis of Date Palm Waste for Adoption in Saudi Arabia. Energies 2021, 14, 6048. [Google Scholar] [CrossRef]
  45. 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]
  46. Valin, S.; Cances, J.; Castelli, P.; Thiery, S.; Dufour, A.; Boissonnet, G.; Spindler, B. Upgrading Biomass Pyrolysis Gas by Conversion of Methane at High Temperature: Experiments and Modelling. Fuel 2009, 88, 834–842. [Google Scholar] [CrossRef]
  47. Lukáč, L.; Kizek, J.; Jablonský, G.; Karakash, Y. Defining the Mathematical Dependencies of NOx and CO Emission Generation after Biomass Combustion in Low-Power Boiler. Civ. Environ. Eng. Rep. 2019, 29, 153–163. [Google Scholar] [CrossRef]
  48. Dhahak, A.; Bounaceur, R.; Le Dreff-Lorimier, C.; Schmidt, G.; Trouve, G.; Battin-Leclerc, F. Development of a Detailed Kinetic Model for the Combustion of Biomass. Fuel 2019, 242, 756–774. [Google Scholar] [CrossRef]
  49. Chang, S.-L.; Zhou, C.Q. Combustion and Thermochemistry. Encycl. Energy 2004, 1, 595–603. [Google Scholar] [CrossRef]
  50. Dupont, C.; Boissonnet, G.; Seiler, J.M.; Gauthier, P.; Schweich, D. Study about the Kinetic Processes of Biomass Steam Gasification. Fuel 2007, 86, 32–40. [Google Scholar] [CrossRef]
  51. Mazaheri, N.; Akbarzadeh, A.H.; Madadian, E.; Lefsrud, M. Systematic Review of Research Guidelines for Numerical Simulation of Biomass Gasification for Bioenergy Production. Energy Convers. Manag. 2019, 183, 671–688. [Google Scholar] [CrossRef]
  52. Gambarotta, A.; Morini, M.; Zubani, A. A Non-Stoichiometric Equilibrium Model for the Simulation of the Biomass Gasification Process. Appl. Energy 2018, 227, 119–127. [Google Scholar] [CrossRef]
  53. Ren, R.; Wang, H.; Feng, X.; You, C. Techno-Economic Analysis of Auto-Thermal Gasification of Municipal Solid Waste with Ash Direct Melting for Hydrogen Production. Energy Convers. Manag. 2023, 292, 117401. [Google Scholar] [CrossRef]
  54. Ferreiro, A.I.; Segurado, R.; Costa, M. Modelling Soot Formation during Biomass Gasification. Renew. Sustain. Energy Rev. 2020, 134, 110380. [Google Scholar] [CrossRef]
  55. Chen, D.; Gao, A.; Cen, K.; Zhang, J.; Cao, X.; Ma, Z. Investigation of Biomass Torrefaction Based on Three Major Components: Hemicellulose, Cellulose, and Lignin. Energy Convers. Manag. 2018, 169, 228–237. [Google Scholar] [CrossRef]
  56. Zhang, J.; Zhang, X. The Thermochemical Conversion of Biomass into Biofuels; Elsevier Ltd.: Amsterdam, The Netherlands, 2019; ISBN 9780081024263. [Google Scholar]
  57. Venturini, P.; Borello, D.; Iossa, C.; Lentini, D.; Rispoli, F. Modeling of Multiphase Combustion and Deposit Formation in a Biomass-Fed Furnace. Energy 2010, 35, 3008–3021. [Google Scholar] [CrossRef]
  58. Ji, W.; Richter, F.; Gollner, M.J.; Deng, S. Autonomous Kinetic Modeling of Biomass Pyrolysis Using Chemical Reaction Neural Networks. Combust. Flame 2022, 240, 111992. [Google Scholar] [CrossRef]
  59. Eke, J.; Onwudili, J.A.; Bridgwater, A.V. Physical Pretreatment of Biogenic-Rich Trommel Fines for Fast Pyrolysis. Waste Manag. 2017, 70, 81–90. [Google Scholar] [CrossRef]
  60. Du, Y.; Ju, T.; Meng, Y.; Han, S.; Jiang, J. Pyrolysis Characteristics of Excavated Waste and Generation Mechanism of Gas Products. J. Clean. Prod. 2022, 370, 133489. [Google Scholar] [CrossRef]
  61. Aboulkas, A.; El Bouadili, A. Thermal Degradation Behaviors of Polyethylene and Polypropylene. Part I: Pyrolysis Kinetics and Mechanisms. Energy Convers. Manag. 2010, 51, 1363–1369. [Google Scholar] [CrossRef]
  62. Till, Z.; Varga, T.; Sója, J.; Miskolczi, N.; Chován, T. Kinetic Modeling of Plastic Waste Pyrolysis in a Laboratory Scale Two-Stage Reactor. Comput. Aided Chem. Eng. 2018, 43, 349–354. [Google Scholar] [CrossRef]
  63. Martínez-Narro, G.; Royston, N.J.; Billsborough, K.L.; Phan, A.N. Kinetic Modelling of Mixed Plastic Waste Pyrolysis. Chem. Thermodyn. Therm. Anal. 2023, 9, 100105. [Google Scholar] [CrossRef]
  64. Abbas-Abadi, M.S.; Haghighi, M.N.; Yeganeh, H.; McDonald, A.G. Evaluation of Pyrolysis Process Parameters on Polypropylene Degradation Products. J. Anal. Appl. Pyrolysis 2014, 109, 272–277. [Google Scholar] [CrossRef]
  65. Ateş, F.; Miskolczi, N.; Borsodi, N. Comparision of Real Waste (MSW and MPW) Pyrolysis in Batch Reactor over Different Catalysts. Part I: Product Yields, Gas and Pyrolysis Oil Properties. Bioresour. Technol. 2013, 133, 443–454. [Google Scholar] [CrossRef] [PubMed]
  66. Acevedo, B.; Fernández, A.M.; Barriocanal, C. Identification of Polymers in Waste Tyre Reinforcing Fibre by Thermal Analysis and Pyrolysis. J. Anal. Appl. Pyrolysis 2015, 111, 224–232. [Google Scholar] [CrossRef]
  67. Policella, M.; Wang, Z.; Burra, K.G.; Gupta, A.K. Characteristics of Syngas from Pyrolysis and CO2-Assisted Gasification of Waste Tires. Appl. Energy 2019, 254, 113678. [Google Scholar] [CrossRef]
  68. Zhang, Y.; Wei, D.; Lv, P.; Liu, Z.; Cheng, T.; Wang, B. Fine particles removal of pyrolysis gasification flue gas from rural domestic waste: Laboratory research, molecular dynamics simulation, and applications. Environ. Res. 2023, 236, 116732. [Google Scholar] [CrossRef]
  69. Porshnov, D.; Ozols, V.; Ansone-Bertina, L.; Burlakovs, J.; Klavins, M. Thermal Decomposition Study of Major Refuse Derived Fuel Components. Energy Procedia 2018, 147, 48–53. [Google Scholar] [CrossRef]
  70. Jannelli, E.; Minutillo, M. Simulation of the Flue Gas Cleaning System of an RDF Incineration Power Plant. Waste Manag. 2007, 27, 684–690. [Google Scholar] [CrossRef] [PubMed]
  71. Çepelioğullar; Haykiri-Açma, H.; Yaman, S. Kinetic Modelling of RDF Pyrolysis: Model-Fitting and Model-Free Approaches. Waste Manag. 2016, 48, 275–284. [Google Scholar] [CrossRef] [PubMed]
  72. Tiwari, S.K.; Bystrzejewski, M.; De Adhikari, A.; Huczko, A.; Wang, N. Methods for the Conversion of Biomass Waste into Value-Added Carbon Nanomaterials: Recent Progress and Applications. Prog. Energy Combust. Sci. 2022, 92, 101023. [Google Scholar] [CrossRef]
  73. Sabogal, O.S.; Valin, S.; Thiery, S.; Salvador, S. Pyrolysis of Solid Waste and Its Components in a Lab Scale Induction-Heating Reactor. Detritus 2021, 15, 107–112. [Google Scholar] [CrossRef]
  74. Xie, J.; Zhong, W.; Jin, B.; Shao, Y.; Liu, H. Simulation on Gasification of Forestry Residues in Fluidized Beds by Eulerian-Lagrangian Approach. Bioresour. Technol. 2012, 121, 36–46. [Google Scholar] [CrossRef] [PubMed]
  75. Radmanesh, R.; Courbariaux, Y.; Chaouki, J.; Guy, C. A Unified Lumped Approach in Kinetic Modeling of Biomass Pyrolysis. Fuel 2006, 85, 1211–1220. [Google Scholar] [CrossRef]
  76. Al-Salem, S.M.; Lettieri, P. Kinetic Study of High Density Polyethylene (HDPE) Pyrolysis. Chem. Eng. Res. Des. 2010, 88, 1599–1606. [Google Scholar] [CrossRef]
  77. Jha, S.; Nanda, S.; Acharya, B.; Dalai, A.K. A Review of Thermochemical Conversion of Waste Biomass to Biofuels. Energies 2022, 15, 6352. [Google Scholar] [CrossRef]
  78. Liu, H.; Alhumade, H.; Elkamel, A. A Combined Scheme of Parallel-Reaction Kinetic Model and Multi-Layer Artificial Neural Network Model on Pyrolysis of Reed Canary. Chem. Eng. Sci. 2023, 281, 119109. [Google Scholar] [CrossRef]
  79. Wang, Q.; Song, H.; Pan, S.; Dong, N.; Wang, X.; Sun, S. Initial Pyrolysis Mechanism and Product Formation of Cellulose: An Experimental and Density Functional Theory(DFT) Study. Sci. Rep. 2020, 10, 3626. [Google Scholar] [CrossRef]
  80. Tang, X.; Xie, Q.; Qiu, R.; Yang, Y. Development of a Relationship between Kinetic Triplets and Heating Rates to Improve Pyrolysis Kinetic Modeling of Polymer. Polym. Degrad. Stab. 2018, 154, 10–26. [Google Scholar] [CrossRef]
  81. González-Vázquez, M.P.; Rubiera, F.; Pevida, C.; Pio, D.T.; Tarelho, L.A.C. Thermodynamic Analysis of Biomass Gasification Using Aspen plus: Comparison of Stoichiometric and Non-Stoichiometric Models. Energies 2021, 14, 189. [Google Scholar] [CrossRef]
  82. Moretti, L.; Arpino, F.; Cortellessa, G.; Di Fraia, S.; Di Palma, M.; Vanoli, L. Reliability of Equilibrium Gasification Models for Selected Biomass Types and Compositions: An Overview. Energies 2022, 15, 61. [Google Scholar] [CrossRef]
  83. Zhang, J.; Zhong, Z.; Zhang, B.; Xue, Z.; Guo, F.; Wang, J. Prediction of Kinetic Parameters of Biomass Pyrolysis Based on the Optimal Mixture Design Method. Clean. Technol. Environ. Policy 2016, 18, 1621–1629. [Google Scholar] [CrossRef]
  84. Xie, T.; Zhao, L.; Yao, Z.; Kang, K.; Jia, J.; Hu, T.; Zhang, X.; Sun, Y.; Huo, L. Co-Pyrolysis of Biomass and Polyethylene: Insights into Characteristics, Kinetic and Evolution Paths of the Reaction Process. Sci. Total. Environ. 2023, 897, 165443. [Google Scholar] [CrossRef] [PubMed]
  85. Mian, I.; Li, X.; Jian, Y.; Dacres, O.D.; Zhong, M.; Liu, J.; Ma, F.; Rahman, N. Kinetic Study of Biomass Pellet Pyrolysis by Using Distributed Activation Energy Model and Coats Redfern Methods and Their Comparison. Bioresour. Technol. 2019, 294, 122099. [Google Scholar] [CrossRef] [PubMed]
  86. Prins, M.J.; Ptasinski, K.J.; Janssen, F.J.J.G. Torrefaction of Wood. Part 1. Weight Loss Kinetics. J. Anal. Appl. Pyrolysis 2006, 77, 28–34. [Google Scholar] [CrossRef]
  87. Hu, S.; Jess, A.; Xu, M. Kinetic Study of Chinese Biomass Slow Pyrolysis: Comparison of Different Kinetic Models. Fuel 2007, 86, 2778–2788. [Google Scholar] [CrossRef]
  88. Zou, J.; Hu, H.; Xue, Y.; Li, C.; Li, Y.; Yellezuome, D.; He, F.; Zhang, X.; Maksudur Rahman, M.; Cai, J. Exploring Kinetic Mechanisms of Biomass Pyrolysis Using Generalized Logistic Mixture Model. Energy Convers. Manag. 2022, 258, 115522. [Google Scholar] [CrossRef]
  89. Torres-Sciancalepore, R.; Asensio, D.; Nassini, D.; Fernandez, A.; Rodriguez, R.; Fouga, G.; Mazza, G. Assessment of the Behavior of Rosa Rubiginosa Seed Waste during Slow Pyrolysis Process towards Complete Recovery: Kinetic Modeling and Product Analysis. Energy Convers. Manag. 2022, 272, 116340. [Google Scholar] [CrossRef]
  90. Wissing, F.; Wirtz, S.; Scherer, V. Simulating Municipal Solid Waste Incineration with a DEM/CFD Method—Influences of Waste Properties, Grate and Furnace Design. Fuel 2017, 206, 638–656. [Google Scholar] [CrossRef]
  91. Kumar, U.; Paul, M.C. CFD Modelling of Biomass Gasification with a Volatile Break-up Approach. Chem. Eng. Sci. 2019, 195, 413–422. [Google Scholar] [CrossRef]
  92. Bieniek, A.; Reinmöller, M.; Küster, F.; Gräbner, M.; Jerzak, W.; Magdziarz, A. Investigation and Modelling of the Pyrolysis Kinetics of Industrial Biomass Wastes. J. Environ. Manag. 2022, 319, 115707. [Google Scholar] [CrossRef]
  93. Debiagi, P.; Nicolai, H.; Han, W.; Janicka, J.; Hasse, C. Machine Learning for Predictive Coal Combustion CFD Simulations—From Detailed Kinetics to HDMR Reduced-Order Models. Fuel 2020, 274, 117720. [Google Scholar] [CrossRef]
  94. Papari, S.; Hawboldt, K. A Review on the Pyrolysis of Woody Biomass to Bio-Oil: Focus on Kinetic Models. Renew. Sustain. Energy Rev. 2015, 52, 1580–1595. [Google Scholar] [CrossRef]
  95. Sommariva, S.; Grana, R.; Maffei, T.; Pierucci, S.; Ranzi, E. A Kinetic Approach to the Mathematical Model of Fixed Bed Gasifiers. Comput. Chem. Eng. 2011, 35, 928–935. [Google Scholar] [CrossRef]
  96. Baruah, D.; Baruah, D.C. Modeling of Biomass Gasification: A Review. Renew. Sustain. Energy Rev. 2014, 39, 806–815. [Google Scholar] [CrossRef]
  97. Ascher, S.; Watson, I.; You, S. Machine Learning Methods for Modelling the Gasification and Pyrolysis of Biomass and Waste. Renew. Sustain. Energy Rev. 2022, 155, 111902. [Google Scholar] [CrossRef]
  98. Grange, N.; Chetehouna, K.; Gascoin, N.; Coppalle, A.; Reynaud, I.; Senave, S. One-Dimensional Pyrolysis of Carbon Based Composite Materials Using FireFOAM. Fire Saf. J. 2018, 97, 66–75. [Google Scholar] [CrossRef]
  99. Díaz González, C.A.; de Oliveira, D.C.; Yepes, D.M.; Pacheco, L.E.; Silva, E.E. Aspen Plus Model of a Downdraft Gasifier for Lignocellulosic Biomass Adjusted by Principal Component Analysis. Energy Convers. Manag. 2023, 296, 117570. [Google Scholar] [CrossRef]
  100. Agu, C.E.; Pfeifer, C.; Eikeland, M.; Tokheim, L.A.; Moldestad, B.M.E. Measurement and Characterization of Biomass Mean Residence Time in an Air-Blown Bubbling Fluidized Bed Gasification Reactor. Fuel 2019, 253, 1414–1423. [Google Scholar] [CrossRef]
  101. Locaspi, A.; Pelucchi, M.; Mehl, M.; Faravelli, T. Towards a Lumped Approach for Solid Plastic Waste Gasification: Polyethylene and Polypropylene Pyrolysis. Waste Manag. 2023, 156, 107–117. [Google Scholar] [CrossRef]
  102. Safavi, A.; Richter, C.; Unnthorsson, R. Revisiting the Reaction Scheme of Slow Pyrolysis of Woody Biomass. Energy 2023, 280, 128123. [Google Scholar] [CrossRef]
  103. Yang, T.; Yuan, G.; Xia, M.; Mu, M.; Chen, S. Kinetic Analysis of the Pyrolysis of Wood/Inorganic Composites under Non-Isothermal Conditions. Eur. J. Wood Wood Prod. 2021, 79, 273–284. [Google Scholar] [CrossRef]
  104. Nawaz, A.; Kumar, P. Pyrolysis Behavior of Low Value Biomass (Sesbania Bispinosa) to Elucidate Its Bioenergy Potential: Kinetic, Thermodynamic and Prediction Modelling Using Artificial Neural Network. Renew. Energy 2022, 200, 257–270. [Google Scholar] [CrossRef]
  105. Kumar, M.; Shukla, S.K.; Upadhyay, S.N.; Mishra, P.K. Analysis of Thermal Degradation of Banana (Musa Balbisiana) Trunk Biomass Waste Using Iso-Conversional Models. Bioresour. Technol. 2020, 310, 123393. [Google Scholar] [CrossRef]
  106. Rasool, T.; Najar, I.; Srivastava, V.C.; Pandey, A. Pyrolysis of Almond (Prunus Amygdalus) Shells: Kinetic Analysis, Modelling, Energy Assessment and Technical Feasibility Studies. Bioresour. Technol. 2021, 337, 125466. [Google Scholar] [CrossRef]
  107. Alsulami, R.A.; El-sayed, S.A.; Eltaher, M.A.; Mohammad, A.; Almitani, K.H.; Mostafa, M.E. Pyrolysis Kinetics and Thermal Degradation Characteristics of Coffee, Date Seed, and Prickly Pear Wastes and Their Blends. Renew. Energy 2023, 216, 119039. [Google Scholar] [CrossRef]
  108. Aboulkas, A.; Makayssi, T.; Bilali, L.; El Harfi, K.; Nadifiyine, M.; Benchanaa, M. Co-Pyrolysis of Oil Shale and Plastics: Influence of Pyrolysis Parameters on the Product Yields. Fuel Process. Technol. 2012, 96, 209–213. [Google Scholar] [CrossRef]
  109. Ranganathan, P.; Gu, S. Computational Fluid Dynamics Modelling of Biomass Fast Pyrolysis in Fluidised Bed Reactors, Focusing Different Kinetic Schemes. Bioresour. Technol. 2016, 213, 333–341. [Google Scholar] [CrossRef] [PubMed]
  110. Qi, J.; Wang, Y.; Hu, M.; Xu, P.; Yuan, H.; Chen, Y. A Reactor Network of Biomass Gasification Process in an Updraft Gasifier Based on the Fully Kinetic Model. Energy 2023, 268, 126642. [Google Scholar] [CrossRef]
  111. Koçer, A.T.; Erarslan, A.; Özçimen, D. Pyrolysis of Aloe Vera Leaf Wastes for Biochar Production: Kinetics and Thermodynamics Analysis. Ind. Crops Prod. 2023, 204, 117354. [Google Scholar] [CrossRef]
  112. Potnuri, R.; Suriapparao, D.V.; Rao, C.S.; Kumar, T.H. Understanding the Role of Modeling and Simulation in Pyrolysis of Biomass and Waste Plastics: A Review. Bioresour. Technol. Rep. 2022, 20, 101221. [Google Scholar] [CrossRef]
  113. EL-Sayed, S.A. Review of Thermal Decomposition, Kinetics Parameters and Evolved Gases during Pyrolysis of Energetic Materials Using Different Techniques. J. Anal. Appl. Pyrolysis 2022, 161, 105364. [Google Scholar] [CrossRef]
  114. Pan, J.; Jiang, H.; Qing, T.; Zhang, J.; Tian, K. Transformation and Kinetics of Chlorine-Containing Products during Pyrolysis of Plastic Wastes. Chemosphere 2021, 284, 131348. [Google Scholar] [CrossRef]
  115. Dorokhov, V.V.; Nyashina, G.S.; Strizhak, P.A. Thermogravimetric, Kinetic Study and Gas Emissions Analysis of the Thermal Decomposition of Waste-Derived Fuels. J. Environ. Sci. 2024, 137, 155–171. [Google Scholar] [CrossRef]
  116. Fakhrhoseini, S.M.; Dastanian, M. Predicting Pyrolysis Products of PE, PP, and PET Using NRTL Activity Coefficient Model. J. Chem. 2013, 2013, 7–9. [Google Scholar] [CrossRef]
  117. Locaspi, A.; Pelucchi, M.; Faravelli, T. Towards a Lumped Approach for Solid Plastic Waste Gasification: Polystyrene Pyrolysis. J. Anal. Appl. Pyrolysis 2023, 171, 105960. [Google Scholar] [CrossRef]
  118. An, S.; Jung, J.C. Kinetic Modeling of Thermal Reactor in Claus Process Using CHEMKIN-PRO Software. Case Stud. Therm. Eng. 2020, 21, 100694. [Google Scholar] [CrossRef]
  119. Alves, J.L.F.; da Silva, J.C.G.; Mumbach, G.D.; Alves, R.F.; Di Domenico, M. Kinetic Triplet and Thermodynamic Parameters of the Pyrolysis Reaction of Invasive Grass Eleusine Indica Biomass: A New Low-Cost Feedstock for Bioenergy Production. Biomass Convers. Biorefinery 2022, 1, 1–17. [Google Scholar] [CrossRef]
  120. Alves, J.L.F.; da Silva, J.C.G.; Mumbach, G.D.; de Sena, R.F.; Machado, R.A.F.; Marangoni, C. Prospection of Catole Coconut (Syagrus cearensis) as a New Bioenergy Feedstock: Insights from Physicochemical Characterization, Pyrolysis Kinetics, and Thermodynamics Parameters. Renew. Energy 2022, 181, 207–218. [Google Scholar] [CrossRef]
  121. Mumbach, G.D.; Alves, J.L.F.; da Silva, J.C.G.; Domenico, M.D.; Marangoni, C.; Machado, R.A.F.; Bolzan, A. Investigation on Prospective Bioenergy from Pyrolysis of Butia Seed Waste Using TGA-FTIR: Assessment of Kinetic Triplet, Thermodynamic Parameters and Evolved Volatiles. Renew. Energy 2022, 191, 238–250. [Google Scholar] [CrossRef]
  122. Mumbach, G.D.; Alves, J.L.F.; da Silva, J.C.G.; Domenico, M.D.; Arias, S.; Pacheco, J.G.A.; Marangoni, C.; Machado, R.A.F.; Bolzan, A. Prospecting Pecan Nutshell Pyrolysis as a Source of Bioenergy and Bio-Based Chemicals Using Multicomponent Kinetic Modeling, Thermodynamic Parameters Estimation, and Py-GC/MS Analysis. Renew. Sustain. Energy Rev. 2022, 153, 111753. [Google Scholar] [CrossRef]
  123. Alves, J.L.F.; da Silva, J.C.G.; Mumbach, G.D.; Alves, R.F.; de Sena, R.F.; Machado, R.A.F.; Marangoni, C. Potential of Macauba Endocarp (Acrocomia Aculeate) for Bioenergy Production: Multi-Component Kinetic Study and Estimation of Thermodynamic Parameters of Activation. Thermochim. Acta 2022, 708, 179134. [Google Scholar] [CrossRef]
  124. Alves, J.L.F.; da Silva, J.C.G.; Mumbach, G.D.; Arias, S.; Pacheco, J.G.A.; Di Domenico, M.; Marangoni, C. Valorization of Royal Palm Tree Agroindustrial Waste via Pyrolysis with a Focus on Physicochemical Properties, Kinetic Triplet, Thermodynamic Parameters, and Volatile Products. Biomass Bioenergy 2023, 177, 106937. [Google Scholar] [CrossRef]
  125. Sharma, A.; Aravind Kumar, A.; Mohanty, B.; Sawarkar, A.N. Critical Insights into Pyrolysis and Co-Pyrolysis of Poplar and Eucalyptus Wood Sawdust: Physico-Chemical Characterization, Kinetic Triplets, Reaction Mechanism, and Thermodynamic Analysis. Renew. Energy 2023, 210, 321–334. [Google Scholar] [CrossRef]
  126. Wang, S.; Lin, H.; Ru, B.; Dai, G.; Wang, X.; Xiao, G.; Luo, Z. Kinetic Modeling of Biomass Components Pyrolysis Using a Sequential and Coupling Method. Fuel 2016, 185, 763–771. [Google Scholar] [CrossRef]
  127. Alves, J.L.F.; da Silva, J.C.G.; Mumbach, G.D.; Alves, R.F.; Di Domenico, M.; Marangoni, C. Physicochemical Properties, Pyrolysis Kinetics, Thermodynamic Parameters of Activation, and Evolved Volatiles of Mango Seed Waste as a Bioenergy Feedstock: A Potential Exploration. Thermochim. Acta 2023, 725, 179519. [Google Scholar] [CrossRef]
  128. Xiong, Q.; Aramideh, S.; Passalacqua, A.; Kong, S.C. BIOTC: An Open-Source CFD Code for Simulating Biomass Fast Pyrolysis. Comput. Phys. Commun. 2014, 185, 1739–1746. [Google Scholar] [CrossRef]
  129. Mahmood, H.; Ramzan, N.; Shakeel, A.; Moniruzzaman, M.; Iqbal, T.; Kazmi, M.A.; Sulaiman, M. Kinetic Modeling and Optimization of Parameters for Biomass Pyrolysis: A Comparison of Different Lignocellulosic Biomass. Energy Sources Part. A Recover. Util. Environ. Eff. 2019, 41, 1690–1700. [Google Scholar] [CrossRef]
  130. Luo, H.; Wang, X.; Krochmalny, K.; Niedzwiecki, L.; Czajka, K.; Pawlak-Kruczek, H.; Wu, X.; Liu, X.; Xiong, Q. Assessments and Analysis of Lumped and Detailed Pyrolysis Kinetics for Biomass Torrefaction with Particle-Scale Modeling. Biomass Bioenergy 2022, 166, 106619. [Google Scholar] [CrossRef]
  131. Di Blasi, C. Modeling Chemical and Physical Processes of Wood and Biomass Pyrolysis. Prog. Energy Combust. Sci. 2008, 34, 47–90. [Google Scholar] [CrossRef]
  132. Nakhaei, M.; Wu, H.; Grévain, D.; Jensen, L.S.; Glarborg, P.; Clausen, S.; Dam-Johansen, K. Experiments and Modeling of Single Plastic Particle Conversion in Suspension. Fuel Process. Technol. 2018, 178, 213–225. [Google Scholar] [CrossRef]
  133. Gao, N.; Chen, C.; Magdziarz, A.; Zhang, L.; Quan, C. Modeling and Simulation of Pine Sawdust Gasification Considering Gas Mixture Reflux. J. Anal. Appl. Pyrolysis 2021, 155, 105094. [Google Scholar] [CrossRef]
  134. Lee, Y.R.; Choi, H.S.; Park, H.C.; Lee, J.E. A Numerical Study on Biomass Fast Pyrolysis Process: A Comparison between Full Lumped Modeling and Hybrid Modeling Combined with CFD. Comput. Chem. Eng. 2015, 82, 202–215. [Google Scholar] [CrossRef]
  135. Akinnawo, O.O.; Nurhafizah, M.D.; Abdullah, N. Pyrolysis Kinetic Study of the Thermal Degradation of Pre-Treated Empty Fruit Bunches. Mater. Today Proc. 2023, 1, 2214. [Google Scholar] [CrossRef]
  136. Matta, J.; Bronson, B.; Gogolek, P.E.G.; Mazerolle, D.; Thibault, J.; Mehrani, P. Comparison of Multi-Component Kinetic Relations on Bubbling Fluidized-Bed Woody Biomass Fast Pyrolysis Reactor Model Performance. Fuel 2017, 210, 625–638. [Google Scholar] [CrossRef]
  137. Che, D.; Li, S.; Yang, W.; Jia, J.; Zheng, N. Application of Numerical Simulation on Biomass Gasification. Energy Procedia 2012, 17, 49–54. [Google Scholar] [CrossRef]
  138. Peters, J.F.; Banks, S.W.; Bridgwater, A.V.; Dufour, J. A Kinetic Reaction Model for Biomass Pyrolysis Processes in Aspen Plus. Appl. Energy 2017, 188, 595–603. [Google Scholar] [CrossRef]
  139. Granados, D.A.; Chejne, F.; Basu, P. A Two Dimensional Model for Torrefaction of Large Biomass Particles. J. Anal. Appl. Pyrolysis 2016, 120, 1–14. [Google Scholar] [CrossRef]
  140. Park, C.; Zahid, U.; Lee, S.; Han, C. Effect of Process Operating Conditions in the Biomass Torrefaction: A Simulation Study Using One-Dimensional Reactor and Process Model. Energy 2015, 79, 127–139. [Google Scholar] [CrossRef]
  141. Bates, R.B.; Ghoniem, A.F. Biomass Torrefaction: Modeling of Volatile and Solid Product Evolution Kinetics. Bioresour. Technol. 2012, 124, 460–469. [Google Scholar] [CrossRef] [PubMed]
  142. Oliveira, M.; Ramos, A.; Monteiro, E.; Rouboa, A. Modeling and Simulation of a Fixed Bed Gasification Process for Thermal Treatment of Municipal Solid Waste and Agricultural Residues. Energy Rep. 2021, 7, 256–269. [Google Scholar] [CrossRef]
  143. Vaishnavi, M.; Vasanth, P.M.; Rajkumar, S.; Gopinath, K.P.; Devarajan, Y. A Critical Review of the Correlative Effect of Process Parameters on Pyrolysis of Plastic Wastes. J. Anal. Appl. Pyrolysis 2023, 170, 105907. [Google Scholar] [CrossRef]
  144. Zhou, C.; Yang, W. Effect of Heat Transfer Model on the Prediction of Refuse-Derived Fuel Pyrolysis Process. Fuel 2015, 142, 46–57. [Google Scholar] [CrossRef]
  145. Aluri, S.; Syed, A.; Flick, D.W.; Muzzy, J.D.; Sievers, C.; Agrawal, P.K. Pyrolysis and Gasification Studies of Model Refuse Derived Fuel (RDF) Using Thermogravimetric Analysis. Fuel Process. Technol. 2018, 179, 154–166. [Google Scholar] [CrossRef]
  146. Zaini, I.N.; García López, C.; Pretz, T.; Yang, W.; Jönsson, P.G. Characterization of Pyrolysis Products of High-Ash Excavated-Waste and Its Char Gasification Reactivity and Kinetics under a Steam Atmosphere. Waste Manag. 2019, 97, 149–163. [Google Scholar] [CrossRef] [PubMed]
  147. Dupont, C.; Chen, L.; Cances, J.; Commandre, J.M.; Cuoci, A.; Pierucci, S.; Ranzi, E. Biomass Pyrolysis: Kinetic Modelling and Experimental Validation under High Temperature and Flash Heating Rate Conditions. J. Anal. Appl. Pyrolysis 2009, 85, 260–267. [Google Scholar] [CrossRef]
  148. Ibrahimoglu, B.; Cucen, A.; Yilmazoglu, M.Z. Numerical Modeling of a Downdraft Plasma Gasification Reactor. Int. J. Hydrog. Energy 2017, 42, 2583–2591. [Google Scholar] [CrossRef]
  149. Sieradzka, M.; Rajca, P.; Zajemska, M.; Mlonka-Mędrala, A.; Magdziarz, A. Prediction of Gaseous Products from Refuse Derived Fuel Pyrolysis Using Chemical Modelling Software—Ansys Chemkin-Pro. J. Clean. Prod. 2020, 248, 119277. [Google Scholar] [CrossRef]
  150. Ślefarski, R.; Jójka, J.; Czyżewski, P.; Gołębiewski, M.; Jankowski, R.; Markowski, J.; Magdziarz, A. Experimental and Numerical-Driven Prediction of Automotive Shredder Residue Pyrolysis Pathways toward Gaseous Products. Energies 2021, 14, 1779. [Google Scholar] [CrossRef]
  151. Zeaiter, J. A Process Study on the Pyrolysis of Waste Polyethylene. Fuel 2014, 133, 276–282. [Google Scholar] [CrossRef]
  152. Adrados, A.; de Marco, I.; Caballero, B.M.; López, A.; Laresgoiti, M.F.; Torres, A. Pyrolysis of Plastic Packaging Waste: A Comparison of Plastic Residuals from Material Recovery Facilities with Simulated Plastic Waste. Waste Manag. 2012, 32, 826–832. [Google Scholar] [CrossRef] [PubMed]
  153. Li, H.; Zhang, S.; Zhao, X.; Eiji, S. Pyrolysis Characteristics and Kinetics of Municipal Solid Waste. Trans. Tianjin Univ. 2005, 11, 353–359. [Google Scholar]
  154. Gautam, R.; Vinu, R. Unraveling the Interactions in Fast Co-Pyrolysis of Microalgae Model Compounds via Pyrolysis-GC/MS and Pyrolysis-FTIR Techniques. React. Chem. Eng. 2019, 4, 278–297. [Google Scholar] [CrossRef]
  155. Chen, R.; Zhang, J.; Lun, L.; Li, Q.; Zhang, Y. Comparative Study on Synergistic Effects in Co-Pyrolysis of Tobacco Stalk with Polymer Wastes: Thermal Behavior, Gas Formation, and Kinetics. Bioresour. Technol. 2019, 292, 121970. [Google Scholar] [CrossRef] [PubMed]
  156. Ma, Y.; Wang, J.; Zhang, Y. Analysis of Pyrolysis Characteristics and Kinetics of Euphausia Superba Shell Waste Using TG-FTIR and Distributed Activation Energy Model. Biomass Convers. Biorefinery 2018, 8, 329–337. [Google Scholar] [CrossRef]
  157. Ma, Z.; Wang, J.; Yang, Y.; Zhang, Y.; Zhao, C.; Yu, Y.; Wang, S. Comparison of the Thermal Degradation Behaviors and Kinetics of Palm Oil Waste under Nitrogen and Air Atmosphere in TGA-FTIR with a Complementary Use of Model-Free and Model-Fitting Approaches. J. Anal. Appl. Pyrolysis 2018, 134, 12–24. [Google Scholar] [CrossRef]
  158. Rajca, P.; Poskart, A.; Chrubasik, M.; Sajdak, M.; Zajemska, M.; Skibiński, A.; Korombel, A. Technological and Economic Aspect of Refuse Derived Fuel Pyrolysis. Renew. Energy 2020, 161, 482–494. [Google Scholar] [CrossRef]
  159. Paraiso, K.; Sauvage, E.; Schuller, S.; Hocine, S.; Lemaitre, V.; Burov, E. Characterization and Modeling of Chemical Reactions Taking Place during the Vitrification of High Level Nuclear Waste. J. Nucl. Mater. 2022, 569, 153878. [Google Scholar] [CrossRef]
  160. Nawaz, A.; Singh, B.; Mishra, R.K.; Kumar, P. Pyrolysis of Low-Value Waste Trapa Natans Peels: An Exploration of Thermal Decomposition Characteristics, Kinetic Behaviour, and Pyrolytic Liquid Product. Sustain. Energy Technol. Assess. 2023, 56, 103128. [Google Scholar] [CrossRef]
  161. Yousef, S.; Eimontas, J.; Striūgas, N.; Abdelnaby, M.A. Pyrolysis Kinetic Behaviour and TG-FTIR-GC–MS Analysis of Coronavirus Face Masks. J. Anal. Appl. Pyrolysis 2021, 156, 105118. [Google Scholar] [CrossRef] [PubMed]
  162. Mesa-Pérez, J.M.; Cortez, L.A.B.; Marín-Mesa, H.R.; Rocha, J.D.; Peláez-Samaniego, M.R.; Cascarosa, E. A Statistical Analysis of the Auto Thermal Fast Pyrolysis of Elephant Grass in Fluidized Bed Reactor Based on Produced Charcoal. Appl. Therm. Eng. 2014, 65, 322–329. [Google Scholar] [CrossRef]
  163. Postawa, K.; Fałtynowicz, H.; Sczygieł, J.; Beran, E.; Kułażski, M. Analyzing the Kinetics of Waste Plant Biomass Pyrolysis via Thermogravimetry Modeling and Semi-Statistical Methods. Bioresour. Technol. 2022, 344, 126181. [Google Scholar] [CrossRef] [PubMed]
  164. Chomsamutr, K.; Jongprasithporn, S. Optimization Parameters of Tool Life Model Using the Taguchi Approach and Response Surface Methodology. Int. J. Comput. Sci. Issues 2012, 9, 120–125. [Google Scholar]
  165. Chen, W.H.; Carrera Uribe, M.; Kwon, E.E.; Lin, K.Y.A.; Park, Y.K.; Ding, L.; Saw, L.H. A Comprehensive Review of Thermoelectric Generation Optimization by Statistical Approach: Taguchi Method, Analysis of Variance (ANOVA), and Response Surface Methodology (RSM). Renew. Sustain. Energy Rev. 2022, 169, 112917. [Google Scholar] [CrossRef]
  166. Mazurek, I.; Skawińska, A.; Sajdak, M. Analysis of Chlorine Forms in Hard Coal and the Impact of Leaching Conditions on Chlorine Removal. J. Energy Inst. 2021, 94, 337–351. [Google Scholar] [CrossRef]
  167. Ajorloo, M.; Ghodrat, M.; Scott, J.; Strezov, V. Modelling and Statistical Analysis of Plastic Biomass Mixture Co-Gasification. Energy 2022, 256, 124638. [Google Scholar] [CrossRef]
  168. Jha, K.K.; Kannan, T.T.M.; Senthilvelan, N. Optimization of Catalytic Pyrolysis Process for Change of Plastic Waste into Fuel. Mater. Today Proc. 2020, 39, 708–711. [Google Scholar] [CrossRef]
  169. Alqarni, A.O.; Nabi, R.A.U.; Althobiani, F.; Naz, M.Y.; Shukrullah, S.; Khawaja, H.A.; Bou-Rabee, M.A.; Gommosani, M.E.; Abdushkour, H.; Irfan, M.; et al. Statistical Optimization of Pyrolysis Process for Thermal Destruction of Plastic Waste Based on Temperature-Dependent Activation Energies and Pre-Exponential Factors. Processes 2022, 10, 1559. [Google Scholar] [CrossRef]
  170. Chen, W.H.; Pratim Biswas, P.; Kwon, E.E.; Park, Y.K.; Rajendran, S.; Gnanasekaran, L.; Chang, J.S. Optimization of the Process Parameters of Catalytic Plastic Pyrolysis for Oil Production Using Design of Experiment Approaches: A Review. Chem. Eng. J. 2023, 471, 144695. [Google Scholar] [CrossRef]
  171. Ali, D.A.; Gadalla, M.A.; Abdelaziz, O.Y.; Hulteberg, C.P.; Ashour, F.H. Co-Gasification of Coal and Biomass Wastes in an Entrained Flow Gasifier: Modelling, Simulation and Integration Opportunities. J. Nat. Gas Sci. Eng. 2017, 37, 126–137. [Google Scholar] [CrossRef]
Figure 1. Hierarchy of treatment of non-hazardous waste [9,10].
Figure 1. Hierarchy of treatment of non-hazardous waste [9,10].
Energies 17 03067 g001
Figure 2. Screenshot of top 20 keywords indicated by the VOSviewer with their number of occurrences and total link strength.
Figure 2. Screenshot of top 20 keywords indicated by the VOSviewer with their number of occurrences and total link strength.
Energies 17 03067 g002
Figure 3. VOSviewer network visualization of keyword co-occurrence.
Figure 3. VOSviewer network visualization of keyword co-occurrence.
Energies 17 03067 g003
Figure 4. Stages of computer simulation [33].
Figure 4. Stages of computer simulation [33].
Energies 17 03067 g004
Figure 5. Motivation, advantages and disadvantages of computer simulations [21,33,39,40].
Figure 5. Motivation, advantages and disadvantages of computer simulations [21,33,39,40].
Energies 17 03067 g005
Figure 6. Simplifications of equilibrium models [99].
Figure 6. Simplifications of equilibrium models [99].
Energies 17 03067 g006
Figure 7. Three-stage pyrolysis reaction diagram [138].
Figure 7. Three-stage pyrolysis reaction diagram [138].
Energies 17 03067 g007
Figure 8. The scheme of chemical reaction kinetics mechanism proposed by Park et al. [140].
Figure 8. The scheme of chemical reaction kinetics mechanism proposed by Park et al. [140].
Energies 17 03067 g008
Figure 9. Model according to Wang et al. [126].
Figure 9. Model according to Wang et al. [126].
Energies 17 03067 g009
Figure 10. Diagram of procedure for modelling chemical composition of pyrolysis gas using Ansys Chemkin-Pro software [149].
Figure 10. Diagram of procedure for modelling chemical composition of pyrolysis gas using Ansys Chemkin-Pro software [149].
Energies 17 03067 g010
Figure 11. The comparison of response surface methodology and Taguchi method [170].
Figure 11. The comparison of response surface methodology and Taguchi method [170].
Energies 17 03067 g011
Figure 12. Comparison of optimization approaches for different methods [165].
Figure 12. Comparison of optimization approaches for different methods [165].
Energies 17 03067 g012
Figure 13. Steps of statistical optimization employing Taguchi, ANNOVA and RSM methods [165].
Figure 13. Steps of statistical optimization employing Taguchi, ANNOVA and RSM methods [165].
Energies 17 03067 g013
Figure 14. Comparison of the composition of bio-oil from beech wood pyrolysis obtained from simulation and literature data [138].
Figure 14. Comparison of the composition of bio-oil from beech wood pyrolysis obtained from simulation and literature data [138].
Energies 17 03067 g014
Figure 15. Fraction yield % [171].
Figure 15. Fraction yield % [171].
Energies 17 03067 g015
Figure 16. Bio-oil compounds [171].
Figure 16. Bio-oil compounds [171].
Energies 17 03067 g016
Figure 17. Char composition [171].
Figure 17. Char composition [171].
Energies 17 03067 g017
Table 1. Use of computer simulations to predict products of thermal conversion of biomass.
Table 1. Use of computer simulations to predict products of thermal conversion of biomass.
ProcessFeedstock TypeKinetic MechanismRes.
Pyrolysis
Single-step and multi-step thermos kinetic study
multi-step mechanism reaction
Biomass: grass Eleusine indica E l e u s i n e   i n d i c a H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [119]
Kinetic triplets C a t o l e   c o c o n u t H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [120]
Butia seed waste (BSW) B S W E x t r a c t i v e s k e V o l a t i l e s + C h a r H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [121]
Pecan nutshell P N S H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [122]
Macauba endocarp (Acrocomia aculeate) M a c a u b a   e n d o c a r p H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [123]
PyrolysisRoyal palm tree waste R o y a l   p a l m t r e e   w a s t e E x t r a c t i v e s k e V o l a t i l e s + C h a r H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [124]
Pyrolysis Poplar and eucalyptus wood sawdustThe kinetic parameters were computed via model-fitting (inflection point and multiple linear regression) and model-free (OFW and KAS) methods.[125]
Thermal conversion Bambusa vulgaris dust (BD) and delonix regia pods peels (DRP)Model-free methods, like Kissinger–Akahira–Sunose (KAS) and Flynn–Wall-Ozawa (FWO), were applied for the determination of kinetic parameters. [126]
PyrolysisSesbania bispinosaThe three-pseudo-component model is made up of hemicellulose, cellulose and lignin.
Artificial neural networks (ANN) are models based on the operation of the human brain.
[104]
CombustionBiomass C e l l u l o s e , h e m i c e l l u l o s e , l i g n i n , b i o m a s s 1 A c t i v e   i n t e r m e d i a t e 2 V o l a t i l e s 4 G a s 3 Y C h a r + 1 Y G a s
(1)
Biomass → Tars
(2)
Biomass → Volatiles
(3)
Biomass → Char
(4)
Tars → Volatiles
(5)
Tars → Char
(6)
Char + O2 → CO2
(7)
Char + CO2 → 2 CO
(8)
Char + H2O → CO + H2
[48]
Pyrolysis M a n g o   s e e d   w a s t e E x t r a c t i v e s k e V o l a t i l e s + C h a r H e m i c e l l u l o s e   k h V o l a t i l e s + C h a r C e l l u l o s e k c V o l a t i l e s + C h a r L i g n i n k l V o l a t i l e s + C h a r [127]
PyrolysisVirgin Energies 17 03067 i001[128]
Pyrolysis three-independent parallel reactions modelBeech wood, Rice huskA three-independent-parallel-reactions model is used to model the kinetics of total devolatilization.[75]
The two-step kinetic model proposed by Koufopanos et al. for lignocellulosic biomass pyrolysisWood sawdust, bagasse, peanut hull, douglas fir bark, and rice husk [129]
Torrefaction processBeechwood
  • lump kinetic scheme;
  • detailed kinetic scheme developed by Andr’es and Ingwald;
  • detailed kinetic scheme developed by Debiagi et al. with a one-dimensional model.
[130]
Pyrolysis Energies 17 03067 i002
One-component mechanism of primary wood pyrolysis proposed by Shafizadeh and Chin.
Multi-component (or multi-stage) mechanisms of wood/biomass pyrolysis.
[131]
PyrolysisBiomassEnergies 17 03067 i003
Comprehensive kinetic models for pyrolysis of cellulose, hemicellulose and lignin.
[126]
CombustionPolyethyleneA Non-isothermal 1D model[132]
Gasification processPine sawdustComprehensive model was developed by Aspen Plus.[133]
PyrolysisWood, grass, and cropsChemical reaction neural networks (CRNN) model[58]
PyrolysisBiomassTwo-stage semi-global mechanism (CFD) Energies 17 03067 i004[134]
Gasification Main kinetic schemes for wood gasification:
(1)
Wood     G a s T a r C h a r → Tar → G a s C h a r
(2)
Cellulose → Cellulose Active → T a r C h a r + G a s T a r G a s
(3)
Virgin   Biomass     V o l i t e + G a s e s C h a r V o l i t e + G a s e s C h a r
[51]
PyrolysisPolietylen, polipropylen, politereftalan etylenuEnergies 17 03067 i005[116]
Pyrolysis Empty fruit bunchA simplified first-order gasification reaction model[135]
Energies 17 03067 i006[136]
TorrefactionWillow Energies 17 03067 i007[86]
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

Skrzyniarz, M.; Sajdak, M.; Biniek-Poskart, A.; Skibiński, A.; Krakowiak, M.; Piotrowski, A.; Krasoń, P.; Zajemska, M. Methods and Validation Techniques of Chemical Kinetics Models in Waste Thermal Conversion Processes. Energies 2024, 17, 3067. https://doi.org/10.3390/en17133067

AMA Style

Skrzyniarz M, Sajdak M, Biniek-Poskart A, Skibiński A, Krakowiak M, Piotrowski A, Krasoń P, Zajemska M. Methods and Validation Techniques of Chemical Kinetics Models in Waste Thermal Conversion Processes. Energies. 2024; 17(13):3067. https://doi.org/10.3390/en17133067

Chicago/Turabian Style

Skrzyniarz, Magdalena, Marcin Sajdak, Anna Biniek-Poskart, Andrzej Skibiński, Marlena Krakowiak, Andrzej Piotrowski, Patrycja Krasoń, and Monika Zajemska. 2024. "Methods and Validation Techniques of Chemical Kinetics Models in Waste Thermal Conversion Processes" Energies 17, no. 13: 3067. https://doi.org/10.3390/en17133067

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