Next Article in Journal
Is Grazing Good for Wet Meadows? Vegetation Changes Caused by White-Backed Cattle
Next Article in Special Issue
Developing and Testing the Air Cooling System of a Combined Climate Control Unit Used in Pig Farming
Previous Article in Journal
Investigating the Effect of Tractor’s Tire Parameters on Soil Compaction Using Statistical and Adaptive Neuro-Fuzzy Inference System (ANFIS) Methods
Previous Article in Special Issue
Optimal Design of Agricultural Mobile Robot Suspension System Based on NSGA-III and TOPSIS
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pyrolysis of Amaranth Inflorescence Wastes: Bioenergy Potential, Biochar and Hydrocarbon Rich Bio-Oil Production

1
Institute of Power Engineering and Advanced Technologies, FRC Kazan Scientific Center, Russian Academy of Sciences, 420111 Kazan, Russia
2
A.E. Arbuzov Institute of Organic and Physical Chemistry, FRC Kazan Scientific Center, Russian Academy of Sciences, 420088 Kazan, Russia
3
Institute of Geology and Petroleum Technologies, Kazan Federal University, 420008 Kazan, Russia
4
Department of Theoretical and Applied Mechanics, Russian University of Transport, 127994 Moscow, Russia
5
Laboratory of Power Supply and Heat Supply, Federal Scientific Agroengineering Center VIM, 109428 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Agriculture 2023, 13(2), 260; https://doi.org/10.3390/agriculture13020260
Submission received: 22 December 2022 / Revised: 13 January 2023 / Accepted: 18 January 2023 / Published: 20 January 2023
(This article belongs to the Special Issue Engineering Innovations in Agriculture)

Abstract

:
Many agro-industrial companies grow amaranth for the subsequent production of amaranth oil, flour, cereals, flakes, and bran. After the grain is extracted, waste in the form of inflorescences remains, which can be used to obtain useful new products. This work investigated the use of pyrolysis to recycle amaranth inflorescence wastes (AIW). Thermochemical conversion experiments in an inert medium were carried out in a laboratory setup at 550 °C and a heating rate of 10 °C/min. It was found that the AIW pyrolysis produced 37.1 wt.% bio-oil, 35.8 wt.% pyrogas and 27.1 wt.% biochar. The oil fraction of the obtained bio-oil contains 41.8% of hydrocarbons. Thermogravimetric analysis of AIW was performed in the temperature range from 40 to 1000 °C at heating rates of 10, 15, and 20 °C/min in argon medium (75 mL/min). The kinetic parameters were determined by the model-free Friedman, Ozawa-Flynn-Wall, and Kissinger-Akahira-Sunose methods. The average activation energy values are in the range of 208.44–216.17 kJ/mol, and they were used to calculate the thermodynamic parameters. The results indicate that the pyrolysis application will allow efficient conversion of AIW into value-added products.

1. Introduction

Amaranth is of great importance for world food security, especially for developing countries in Africa and Asia [1,2]. These plants are pseudocereals that were very important for ancient civilizations [3]. Currently, interest in this culture is growing for a number of reasons. First, amaranth can grow in a wide range of weather conditions and is drought-tolerant. Secondly, the growing demand for a healthy diet encourages the use of this plant. Amaranth is recognized as a rich and inexpensive source of dietary fiber, minerals, vitamins, proteins, and antioxidants [2,4]. One of the most common cereal species is Amaranthus cruentus [5]. After the grain is extracted, waste in the form of inflorescences remains, which can be used to obtain useful new products.
Pyrolysis is a technology widely used for waste disposal [6,7,8]. The uniqueness of this process lies in the simultaneous production of gaseous, liquid and solid products. The sphere of use of these products is quite wide, including the chemical industry [9], energy [10,11], and agro-industrial complex [12,13]. The process of biomass pyrolysis can be carried out in a decentralized manner, which is especially important for agriculture [14]. It is necessary to assess the energy potential of the resulting pyrolysis products in order to assess the possibility of creating a local non-volatile enterprise. It should be noted that biooil is of particular interest since its composition is very complex and largely depends on the feedstock [15]. It is important to find such a source of biomass, which initially, without the use of catalysts during thermochemical conversion and joint pyrolysis with polymers, contains a significant amount of hydrocarbons.
Currently, thermogravimetric analysis (TGA) is used for a detailed assessment of the pyrolysis process [16,17,18]. TGA data are used to study the kinetics of the thermochemical conversion. This allows a comprehensive study of pyrolysis reactions, revealing the characteristic mechanism and predicting the degree of complexity of the reaction, which is necessary for designing pyrolysis apparatuses and evaluating the possibilities of using products [19]. The kinetics of the amaranth inflorescence wastes (AIW) pyrolysis process was studied using the model-free methods of Friedman, OFW (Ozawa-Flynn-Wall) and KAS (Kissinger-Akahira-Sunose), since they have shown themselves to be effective in assessing the kinetics of biomass pyrolysis reactions [6,20,21,22,23,24,25,26]. Currently, numerous studies are being carried out on the use of biomass as a raw material for thermal decomposition [27,28]; however, there are few works on the pyrolysis of amaranth [8,29,30,31], and no study has yet been reported on the thermal decomposition characteristics of the inflorescences of this plant.
This study is aimed at solving the following problems: (a) determining the material balance of the pyrolysis process of a new type of plant waste; (b) study of the composition and quality of the resulting pyrolysis products to assess their subsequent use; (c) analysis of the features of thermal decomposition of waste according to TGA data at heating rates of 10, 15 and 20 °C/min in an inert atmosphere; (d) determination of kinetic triplets for the main stage of pyrolysis—isolation of volatile components using model-free methods; and (e) determination of thermodynamic functions for subsequent design, optimization and scaling of the parameters of the pyrolysis reactor. Thus, the cultivation of amaranth and the subsequent pyrolysis of the remaining waste will improve not only food, but also energy security, which is especially important for countries with adverse climatic conditions.

2. Materials and Methods

2.1. Amaranth Inflorescence Wastes

AIW samples were taken from a farm after harvest (Russia). Inflorescences were dried at room temperature. All samples of AIW were manually cut into small pieces with the help of blades and then finely powdered using a mixer-cum-grinder. All powdered samples were kept in airtight containers for use in further experiments.

2.2. Physicochemical Characterization

The proximate analysis of all samples was performed to estimate the volatile matter, ash content, moisture, and fixed carbon using appropriate ASTM protocols (E1755-01, 2020; ASTM E1756-08, 2020; E871-82, 2019; E872-82, 2019; D1762-84, 2021). The percentages of carbon, hydrogen, nitrogen, and sulfur were determined using the CHNS analyzer (Euro EA 3000, Eurovector, S.p.A., Milan, Italy) and oxygen was calculated by difference. The higher heating value (HHV) was calculated according to the formula presented in the work [32]. The content of macro- and microelements was studied using the energy-dispersive fluorescence X-ray spectrometer (EDX-800HS2, Shimadzu, Kyoto, Japan) by a semi-quantitative method. Gas chromatography–mass spectrometry of the pyrolysis liquid were carried out on spectrometer (GCMS-QP2010, Shimadzu, Japan) on HP-5MS column (0.25 μm, 30 m). The evolved gas was analyzed by a gas chromatograph Chromatec-Crystal 5000.2 (Chromatec, Yoshkar-Ola, Russia) using GOST 32507−2013 and ASTM D 5134-98, 2008. The gas samples were delivered to the given machine from the autoclave’s gas output through special heat-resistant tubing. The gas separation was carried out in capillary column with a length of 30 m and two absorption chambers. Chromatography was run in following temperature mode: 90 degrees for 4 min, from 90 °C to 250 °C with the heating rate of 10 °C/min. The gas carrier was helium and the stream velocity was 2.5 mL/min.

2.3. Pyrolysis Experimental Procedure

Pyrolysis was carried out on a laboratory setup described in [8]. The initial temperature was 25 °C, the heating rate was 10 °C/min, and the temperature of the pyrolysis process itself was 550 °C. The material balance was determined according to the method presented in [8,33]. Each experiment was repeated at least three times.

2.4. Thermogravimetric Analysis (TGA)

A thermogravimetric analyzer (STA 449 F1 Jupiter, Netzsch, Selb, Germany) was applied to record the mass loss as a function of temperature during the pyrolysis of AIW. The sample was placed in a crucible and heated from 40 to 1000 °C at three different heating rates (10, 15, 20 °C/min) with argon flowing at 75 mL/min. To ensure the repeatability of the experiment with an error of 1.5%, the experimental conditions were repeated at least three times. The results showed that the TG and DTG curves were almost identical, which consequently gave very low standard deviations for the calculated kinetic parameters obtained. The results presented here are a set of those experiments that satisfy the above conditions.

2.5. Kinetic Analysis

The mechanism of pyrolysis is characterized by a rather complex set of competitive and parallel reactions, which is also complicated by the variability of the lignocellulosic composition of the biomass [34]. The global pyrolysis reaction is expressed by the following equation [35]:
Biomass   solid k T Volatiles   condencable + noncondencable + Char ,
where k T is reaction rate constant, which is expressed by the Arrhenius equation:
k T = A e E α / R T ,
where Eα is the activation energy (kJ/mol); T is temperature (K); R is the universal gas constant (8.314 J/mol∙K); and A is the pre-exponential factor (1/s).
The biomass conversion rate α is defined as the mass fraction of the degraded sample. It can be calculated for each point of TGA according to the equation [36,37,38]:
α = m 0 m m 0 m f ,
where m0 is the initial sample weight of the sample (mg); m is actual weight to each point of analysis (mg); and mf is the final weight of the sample after pyrolysis (mg).
Generalized fundamental expression for non-isothermal TGA experiments at linear heating rate:
β = d T d t = d T d α d α d t ,
Or
d α d t = A β exp E α R T f α ,
where f(α) is function of conversion.
The analytical form of the function f(α) depends on the thermal decomposition mechanism. Integration of Equation (5) makes it possible to analyze the kinetic data obtained by the TGA method. Integration can be performed using model-free (isoconversion) methods [38,39], which estimate the activation energy () when changing the degree of conversion α. These methods are also called “multi-curve” since they require the use of several kinetic curves for analysis [40].

2.6. Model-Free Methods

In this work, the modeless methods of Friedman, OFW, and KAS were used to analyze the kinetic parameters [6,20].
The Friedman’s method is a differential method which is expressed by the equation:
ln β i d α d T α , i = ln A α f α E α R T ,
where the subscript i is given heating rate value, and subscript α is given conversion degree.
The OFW method is an integral method which is expressed by the equation:
ln β i = ln A α E α R g α 5.331 1.052 E α R T α i ,
The KAS method used for kinetic determination is given in equation:
ln β i T α i 2 = ln A α R E α   g α E α R T α i ,
where g(α) is constant with given conversion value.

2.7. Reaction Model Determination for AIW Pyrolysis

The master-plot method is used to predict solid state mechanisms in the thermal decomposition of biomass. The master graph is built either in a differential or in a differential-integral form [37]. Various models are fitted to solid-phase kinetic data based on such reaction mechanisms as nucleation, geometric shape, diffusion, and reaction order [19,40]. The theoretical master plots do not depend on the heating rate, but strictly depend on the kinetic model used to model the reaction [41]. To construct a differential graph, a comparison is used at the control point α = 0.5 [42].
d α d θ d α d θ 0.5 = f α f 0.5 ,
where f α f 0.5 – theoretically determined from the function, the expressions for which are given in [43]; θ denotes the reaction time taken to attain a particular α at infinite temperature. The left side of expression (9) is the experimental curve calculated using the following equation:
d α d θ d α d θ 0.5 = d α d t d α d t 0.5 exp E R T exp E R T 0.5 ,
where T 0.5 is the reaction temperature at α = 0.5.

2.8. Thermodynamic Parameters

Estimating thermodynamic parameters is a useful tool for understanding biomass pyrolysis, determining the feasibility of a thermal decomposition process, and calculating energy performance. Enthalpy change ∆H (kJ/mol), Gibbs free energy ∆G (kJ/mol), and entropy change ∆S (J/mol∙K) were calculated according to the equations derived from the activation complex theory (Eyring Theory) using the following formulas [44,45,46]:
Δ H = E α R T p e a k ,
Δ G = E α + R T p e a k ln K B T p e a k h A ,
Δ S = Δ H Δ G T p e a k ,
where T p e a k is the temperature corresponding to the maximum mass loss rate, °C; K B is Boltzmann constant (1.38 ∙ 10−23 J/K); h is Planck’s constant (6.626 ∙ 10−34 J∙s).

3. Results and Discussion

3.1. Results of Proximate and Ultimate Analyses

To assess the possibility of using AIW as a bioenergy raw material, the main physical and chemical characteristics were considered (Table 1).
Humidity and ash content in the AIW sample corresponds to the range of values typical for commercial biomass fuels (humidity up to 25.6 wt.%, ash content up to 9.8 wt.%) [47]. The test sample has a high content of volatile substances; therefore, it is suitable for various thermochemical processes due to its high flammability. The obtained value of volatile substances is comparable with the values obtained for other agricultural wastes suitable for energy use [47,48]. In addition, this means that the AIW sample is more reactive than traditional energy sources such as coal. The HHV of the sample corresponds to the commercial fuel olive stone (17.88 MJ/kg), energy crops–thistle (17.75 MJ/kg) [47], as well as such biomass as: apple tree branches (17.82 MJ/kg), feijoa leaves (17.81 MJ/kg), hazelnut tree leaves (17.87 MJ/kg), kiwi branches (17.81 MJ/kg), and olive stone (17.88 MJ/kg) [49].

3.2. Pyrolysis Products Yields and Their Quality

The pyrolysis products of AIW are shown in Figure 1. The presented values are consistent with the data obtained from the pyrolysis of rice husks [34], switchgrass [50], algal waste [51], and poultry litter [52]. The maximum mass fraction of 37.1 wt.% is characteristic of the pyrolysis liquid. In connection with the subsequent use of pyrolysis liquid for energy purposes, it was separated into oil and aqueous fractions. It is important to use a homogeneous fuel to ensure timely ignition, as well as efficient atomization in the combustion zone and maintaining flame stability in combustion devices [53].
The aqueous fraction of the pyrolysis liquid contains 85.72% water, 10.4% acetic acid, and 3.88% unidentified components. Water is the main component in the liquid, which is explained by the humidity of the AIW sample, dehydration reactions at temperatures below 550 °C, and the occurrence of secondary cracking reactions of oxygen-containing macromolecular compounds at high temperatures [54]. The oil fraction has a diverse and rich composition. Approximately 70.85% of the relative content of the total peak area was identified (GC-MS analysis). The identified compounds were classified into the following main chemical categories: hydrocarbons, phenols, alcohols, ketones, ethers, and N-containing heterocycles. Components with a peak area ≥ 1% are presented in Table 2. Saturated hydrocarbons tetratetracontane and tetracontane are present in large quantities. It is known that the oil fraction from red amaranth seeds is a rich source of squalene, so the content of 2,6,10,14,18,22-Tetracosahexaene, 2,6,10,15,19,23-hexamethyl-, (all-E) equals 5.44% [55]. All identified compounds (including peak area ≤ 1%) are grouped and shown in Figure 2. The oil fraction contains 41.8% hydrocarbons, which characterizes it as a high-quality fuel.
It should also be noted that there are no organic acids in the oil fraction; they are present only in the composition of esters. Accordingly, the pH value is high and the liquid is characterized by an alkaline reaction, which is also important for the design of power plants.
The concentrations of the detected pyro-gas components, converted to nitrogen-free composition, are shown in Table 3. It was found that the predominant components in AIW pyrolysis are CO2 and CO. The total concentration of these gases reaches 94.44%. The combustible part of the pyrolysis gas includes 52.7% of the components, which is consistent with the data of other authors [52,56,57].
The main physicochemical characteristics of AIW biochar (Table 4) correspond to biochars obtained by pyrolysis (process temperature 500 °C) of different biomass [58]. The elemental composition of AIW biochar is typical, in which the content of carbon is in the range of 50–87.2%, hydrogen 0.7–4.5%, nitrogen 0.08–6.94%, and oxygen 6–30% [59]. Figure 3 shows the microelement composition of the ash of the solid carbonaceous residue. The predominant components of the ash were K and Ca, and their total content was 81.8% of the total mass.
Thus, the studied biochar can serve as a direct source of potassium, which is the most important element—a biophile, the removal of which with the harvest of agricultural crops is always greater than that of phosphorus and nitrogen. An analysis of the literature showed that elevated values of K, Mg, and Ca in the solid pyrolysis product make it possible to use it for liming and neutralizing acidic soils [60,61].

3.3. Thermal Degradation Analysis

The results of pyrolysis of AIW samples at heating rates of 10, 15, and 20 °C/min in an argon atmosphere are shown in Figure 4. The TG curves are the change in weight loss with temperature, and the DTG curves are the rate of weight loss with temperature. According to the shape of the curves, it can be judged that the thermal degradation of the studied AIW sample occurs similarly to the general trend of biomass pyrolysis.
Based on the analysis of the obtained TG data, the AIW pyrolysis process can be divided into 3 main stages (Table 5).
The first stage in the temperature range from 40 °C to 190 °C corresponded to the process of evaporation of physically bound moisture from samples of AIW. It is also possible to release light volatile components at this stage [46]. The average weight loss at this stage was 9.25 wt.% for three heating rates (Table 6). The first stage has a small peak characterized by an endothermic reaction, which is associated with the absorption of heat in the process of moisture evaporation [44].
The main stage, corresponding to the main pyrolysis, occurred in the temperature range from 190 °C to the temperature range of 530–560 °C for three heating rates and was accompanied by the main loss of organic matter mass. During this stage, there was an active decomposition of the biomass components and the release of volatile substances associated with the thermal destruction of hemicellulose, cellulose, and lignin [62,63]. The average weight loss during the release of volatiles was 59.63 wt.%. As can be seen from Figure 4, rapid weight loss begins above a temperature of 190 °C, which is associated with the rapid breakdown of thermally unstable components of hemicellulose and extractives [37,64]. Hemicellulose consists of short chain heteropolysaccharides and is an amorphous and branched structure [8,19,33,37,39,41,65,66]. Furthermore, with an increase in the pyrolysis temperature, cellulose is involved in the degradation process, which is characterized by a higher decomposition temperature (315–400 °C) due to the presence of a long polymer of glucose units and a large number of hydrogen bonds in its composition [67]. Cellulose, due to its chemical structure, is more resistant to thermal degradation; its decomposition is typical for the temperature range of 270–350 °C [68,69].
On the DTG-curves (Figure 4) at the stage of devolatilization, one can note the maximum temperature peak, which has values of 317.7, 322.6, and 328.5 °C for the three heating rates. This peak is characterized by an endothermic reaction. In addition, a small temperature exothermic peak is found at 402.7–422.1 °C, which can be associated with the beginning of the decomposition of lignin in the test sample. The literature data indicate that the onset of lignin decomposition for various types of biomass occurs in the temperature range of 280–550 °C [16]. The mechanism of lignin pyrolysis is more complex than that of cellulose and hemicellulose; it includes reactions of free radicals [70,71]. Due to the fact that lignin has the highest thermal stability, it decomposes slowly throughout the thermal degradation up to a temperature of 900 °C [16].
The third stage, which is typical for the temperature range of 529.5 °C and up to 1000 °C for three heating rates, is associated with the process of degradation of char and minerals. At this stage, after the completion of the release of volatiles and the main thermal destruction, the process of enrichment with carbon and the formation of the structure of carbonaceous matter continue. Although small, inorganic minerals in biomass can have a significant effect on the pyrolysis process. In this regard, the process of thermal degradation of mineral components is primarily associated with the decomposition of CaCO3 in the temperature range from 780 to 1000 °C [72]. In addition, pyrolysis products can interact with inorganic elements in the residual carbonaceous matter [73]. In this case, the mineral components act as catalysts in the reactions of gas formation from pyrolysis products [72]. The residual fraction as a result of AIW pyrolysis was 25.5 wt.% for the three heating rates. As a result of the experiments, it was revealed that the nature of the TG and DTG curves of the studied samples is similar to the biomass of herbaceous plants, which were reported in [16,17,74].

3.4. Kinetic Analysis

In this work, a kinetic analysis was carried out for the main stage of pyrolysis—devolatilization—since at this stage, the maximum mass loss occurs [75]. AIW kinetic parameters were determined using three model-free methods: Friedman, KAS, and OFW, based on TGA data. Figure 5 shows the results of linear regression in the range of conversions from 0.1 to 0.9 of the kinetic analysis of the total thermal decomposition reactions of the AIW samples. Straight line slope data obtained from each model-free method were used to calculate the values presented in Table 7.
The dynamics of values obtained by the Friedman, KAS, and OFW methods highlight the complexity of the AIW sample kinetics. gradually increases until reaching its maximum at a conversion rate of 0.8 for the OFW and KAS methods, and α conversion rate of 0.7 for the Friedman method. A similar trend in values was found during pyrolysis of such biomass as bark of Ficus natalensis [7], water hyacinth [76], elephant grass [77], and mustard stalk [78].
The pre-exponential factor A characterizes the frequency of collisions of reacting molecules. This indicator makes it possible to explain the chemistry of reactions, which is important for optimizing the pyrolysis process [36]. Almost all obtained A values are in the range from 104 to 109, which indicates a low reactivity of the test sample and the occurrence of a surface reaction, as well as a tight junctional complex (closed complex) [36,39].
The values are in the range of 152.52–291.94 kJ/mol (Friedman), 156.78–265.25 kJ/mol (KAS), and 157.00–265.43 kJ/mol (OFW). The value of shows a measure of the minimum energy required to start a chemical reaction, as well as a potential measure of reactivity [79,80]. According to the literature data, the KAS and OFW methods are less accurate than the Friedman method [39,79], since it does not contain assumptions and approximations [39,81]. It should be noted that the Eα values calculated by the KAS, OFW, and Friedman methods for the AIW sample agree with each other. The average value of obtained by the Friedman method is only 3.7% higher than that calculated by the OFW and KAS methods. Comparative analysis of values for different types of biomass is presented in Table 8.
Figure 6 compares the theoretical differential plots of f(α))/f(0.5) versus α with the experimental plot of (dα/dθ)/(dα/dθ) 0.5 versus α for a heating rate of 10 °C/min to draw a conclusion about the reaction mechanism of solid-phase pyrolysis.
In the range of α from 0.1 to 0.5, the AIW degradation mechanism refers to the one-dimensional diffusion (D1) process, i.e., heat transfer in the sample occurs by diffusion. When α values are greater than 0.5, the AIW degradation mechanism tends to random nucleation with one nucleus in a single particle (F1). In the range of α from 0.7 to 0.9, the mechanism is reduced to random nucleation with two nuclei in the individual particle (F2). The F1 and F2 degradation mechanisms are initiated from random points that act as growth centers for the development of the degradation reaction [87]. Similar results were obtained for other types of biomass [88]. A slight discrepancy between the experimental curves of the master plot for the studied AIW samples can be explained by the deviation of the ideal conditions adopted in the kinetic models from the actual pyrolysis conditions.

3.5. Thermodynamic Analysis

To design, optimize, and scale the parameters of the pyrolysis reactor, it is also necessary to know the thermodynamic properties of the feedstock used. Thermodynamic parameters were determined for a heating rate of 10 °C/min (Figure 7). In biomass pyrolysis, ΔH is the total energy required for biomass decomposition into solid, liquid, and gaseous products [45,89,90]. The ΔH values for the studied AIW sample were in the range of 152.28–287.03 kJ/mol according to the Friedman method, 151.86–260.34 kJ/mol according to the KAS method, and 152.05–260.52 kJ/mol according to the FWO method. Positive values of ΔH indicate the endothermic nature of biomass pyrolysis, which implies the need for energy from an external heat source [6]. The difference between the average values of and ΔH is insignificant, approximately 5 kJ/mol (for all methods), which indicates that the studied AIW sample is suitable for pyrolysis [43,87,88,90,91,92,93,94].
The change in ΔG makes it possible to judge the energy available in biomass [6,49,95,96]. The ΔG values for the studied AIW sample ranged from 147.90 to 150.78 kJ/mol by the Friedman method, 148.38–150.19 kJ/mol by the KAS method, and 148.38–150.18 kJ/mol by the OFW method. For most known biomasses and their mixtures, the ΔG values are positive [87,90,95,96,97,98]. The resulting average value of ΔG is 149.61 kJ/mol of the AIW sample. It is comparable to the ΔG values for pseudo-hemicelluloses of cocoa shell (143.19 kJ/mol) [99], torrefied biomass of Acacia nilotica T-250 (159.97 kJ/mol) [100], and red macroalgae Gelidium floridanum for stage 1 (147.25 kJ/mol) [44], but higher than mustard stalk (128 kJ/mol) [78]. The data obtained indicate the high energy potential of AIW.
Entropy is a function of the state of a thermodynamic system, which characterizes the direction of spontaneous processes and is a measure of their irreversibility. The change in ΔS serves as a measure of the change in the order of a thermodynamic system. The entropy of the system is the higher the greater the degree of disorder of the system. Thus, if the process goes in the direction of increasing the disorder of the system, then ΔS is a positive value. To increase the degree of order in the system, it is necessary to expend energy [90,93]. The ΔS values of amaranth samples range from 2.15 to 235.45 J/mol∙K by the Friedman method and 1.42-189.78 J/mol∙K by the KAS and OFW methods. Throughout the conversion process, the ΔS values were positive for the three model-free methods, indicating a high reactivity of the biomass and a rapid formation of the activated complex. It should be noted that the degree of disorder in the resulting products was quite high, and this is typical of the pyrolysis process [48,100]. The mean ΔS of the AIW sample was 105.41 J/mol∙K by the Friedman method, 91.15 J/mol∙K by the KAS method, and 91.46 J/mol∙K by the OFW method. The ΔS value is comparable with the values obtained for mixtures of sugarcane bagasse, water hyacinth Eichhornia crassipes and yellow oleander Thevetia Peruviana [101].

4. Conclusions

In this work, a study was made of the pyrolysis of a new type of plant waste using TGA and experiments in a laboratory installation for thermochemical processing. The physicochemical parameters of the studied raw materials correspond to the range of values typical for commercial biomass fuels. The test sample has a high content of volatile substances and high reactivity. The maximum specific gravity in the pyrolysis products of 37.1% corresponds to the pyrolysis liquid. The maximum mass fraction in the pyrolysis products of 37.1 wt.% corresponds to the pyrolysis liquid. At the same time, the oil fraction contains 41.8% hydrocarbons, which characterizes it as a high-quality fuel. Analysis of the features of thermal decomposition of waste was determined at heating rates of 10, 15, and 20 °C/min in an inert atmosphere. The main stage of thermochemical degradation is devolatilization. The kinetic parameters for this stage were determined using the model-free methods of Friedman, OFW, and KAS. The one-dimensional diffusion model (D1), then random nucleation with two nuclei in the individual particle (F1), and random nucleation with two nuclei on the individual particle (F2) were recommended to describe the mechanism of AIW thermal destruction. The average activation energy values are in the range of 208.44–216.17 kJ/mol, and they were used to calculate the thermodynamic parameters. The results indicate that the pyrolysis application will allow the efficient conversion of AIW into value-added products.

Author Contributions

Conceptualization, J.K.; methodology, J.K.; software, S.T. and V.P.; validation, V.B. and V.P.; formal analysis, S.I. and F.A.; investigation, S.T., S.I., K.B. and F.A.; resources, J.K., V.B. and V.P.; data curation, J.K., S.T., V.B. and V.P.; writing—original draft preparation, S.T., S.I., K.B. and F.A.; writing—review and editing, J.K. and V.B.; visualization, S.T.; supervision, V.B. and V.P.; project administration, J.K.; funding acquisition, V.B. and V.P;. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Emmanuel, O.C.; Babalola, O.O. Amaranth production and consumption in South Africa: The challenges of sustainability for food and nutrition security. Int. J. Agric. Sustain. 2022, 20, 449–460. [Google Scholar] [CrossRef]
  2. Sulaiman, M.I.; Andini, R. Potential of Amaranth in Alleviating Malnutrition in Indonesia. In Nutritional Value of Amaranth; Waisundara, Y., Ed.; IntechOpen: London, UK, 2020. [Google Scholar]
  3. Schmidt, D.; Verruma-Bernardi, M.R.; Forti, V.A.; Borges, M.T.M.R. Quinoa and Amaranth as Functional Foods: A Review. Food Rev. Int. 2021, 37, 1–20. [Google Scholar] [CrossRef]
  4. Adhikary, D.; Khatri-Chhetri, U.; Slaski, J. Amaranth: An Ancient and High-Quality Wholesome Crop. In Nutritional Value of Amaranth; Waisundara, Y., Ed.; IntechOpen: London, UK, 2020. [Google Scholar]
  5. Aderibigbe, O.R.; Ezekiel, O.O.; Owolade, S.O.; Korese, J.K.; Sturm, B.; Hensel, O. Exploring the potentials of underutilized grain amaranth (Amaranthus spp.) along the value chain for food and nutrition security: A review. Crit. Rev. Food Sci. Nutr. 2022, 62, 656–669. [Google Scholar] [CrossRef] [PubMed]
  6. Wen, Y.; Shi, Z.; Wang, S.; Mu, W.; Jönsson, P.G.; Yang, W. Pyrolysis of raw and anaerobically digested organic fractions of municipal solid waste: Kinetics, thermodynamics, and product characterization. Chem. Eng. J. 2021, 415, 129064. [Google Scholar] [CrossRef]
  7. Farooq, A.; Ashraf, M.; Aslam, Z.; Anwar, A.; Jiang, S.; Farooq, A.; Liu, L. Pyrolytic conversion of a novel biomass Ficus na-talensis barkcloth: Physiochemical and thermo-kinetic analysis. Biomass Conv. Bioref. 2021, 11. [Google Scholar] [CrossRef]
  8. Karaeva, J.V.; Timofeeva, S.S.; Islamova, S.I.; Gerasimov, A.V. Pyrolysis kinetics of new bioenergy feedstock from anaerobic digestate of agro-waste by thermogravimetric analysis. J. Environ. Chem. Eng. 2022, 10, 107850. [Google Scholar] [CrossRef]
  9. Figueirêdo, M.B.; Hita, I.; Deuss, P.J.; Venderbosch, R.H.; Heeres, H.J. Pyrolytic lignin: A promising biorefinery feedstock for the production of fuels and valuable chemicals. Green Chem. 2022, 24, 4680–4702. [Google Scholar] [CrossRef]
  10. Yang, Q.; Zhou, H.; Bartocci, P.; Fantozzi, F.; Mašek, O.; Agblevor, F.A.; Wei, Z.; Yang, H.; Chen, H.; Lu, X.; et al. Prospective contributions of biomass pyrolysis to China’s 2050 carbon reduction and renewable energy goals. Nat. Commun. 2021, 12, 1698. [Google Scholar]
  11. Mlonka-Mędrala, A.; Evangelopoulos, P.; Sieradzka, M.; Zajemska, M.; Magdziarz, A. Pyrolysis of agricultural waste biomass towards production of gas fuel and high-quality char: Experimental and numerical investigations. Fuel 2021, 296, 120611. [Google Scholar] [CrossRef]
  12. Vuppaladadiyam, A.K.; Vuppaladadiyam, S.S.V.; Sahoo, A.; Murugavelh, S.; Anthony, E.; Bhaskar, T.; Zheng, Y.; Zhao, M.; Duan, H.; Zhao, Y.; et al. Bio-oil and biochar from the pyrolytic conversion of biomass: A current and future perspective on the trade-off between economic, environmental, and technical indicators. Sci. Total. Environ. 2023, 857, 159155. [Google Scholar] [CrossRef]
  13. Lobzenko, I.; Burachevskaya, M.; Zamulina, I.; Barakhov, A.; Bauer, T.; Mandzhieva, S.; Sushkova, S.; Minkina, T.; Tereschenko, A.; Kalinichenko, V.; et al. Development of a Unique Technology for the Pyrolysis of Rice Husk Biochar for Promising Heavy Metal Remediation. Agriculture 2022, 12, 1689. [Google Scholar] [CrossRef]
  14. Ansari, K.B.; Gaikar, V.G. Investigating production of hydrocarbon rich bio-oil from grassy biomass using vacuum pyrolysis coupled with online deoxygenation of volatile products over metallic iron. Renew. Energy 2019, 130, 305–318. [Google Scholar] [CrossRef]
  15. Giorcelli, M.; Das, O.; Sas, G.; Försth, M.; Bartoli, M. A Review of Bio-Oil Production through Microwave-Assisted Pyrolysis. Processes 2021, 9, 561. [Google Scholar] [CrossRef]
  16. Chen, H.; Liu, Z.; Chen, X.; Chen, Y.; Dong, Z.; Wang, X.; Yang, H. Comparative pyrolysis behaviors of stalk, wood and shell biomass: Correlation of cellulose crystallinity and reaction kinetics. Bioresour. Technol. 2020, 310, 123498. [Google Scholar] [CrossRef] [PubMed]
  17. Nour, M.; Amer, M.; Elwardany, A.; Attia, A.; Li, X.; Nada, S. Pyrolysis, kinetics, and structural analyses of agricultural residues in Egypt: For future assessment of their energy potential. Clean. Eng. Technol. 2021, 2, 100080. [Google Scholar] [CrossRef]
  18. Pu, X.; Wei, M.; Chen, X.; Wang, L.; Deng, L. Thermal Decomposition Characteristics and Kinetic Analysis of Chicken Manure in Various Atmospheres. Agriculture 2022, 12, 607. [Google Scholar] [CrossRef]
  19. Tian, B.; Wang, X.; Zhao, W.; Xu, L.; Bai, L. Pyrolysis behaviors, kinetics and gaseous product evolutions of two typical bi-omass wastes. Catal. Today 2021, 374, 77–85. [Google Scholar] [CrossRef]
  20. Ahmad, M.S.; Mehmood, M.A.; Taqvi, S.T.H.; Elkamel, A.; Liu, C.-G.; Xu, J.; Rahimuddin, S.A.; Gull, M. Pyrolysis, kinetics analysis, thermodynamics parameters and reaction mechanism of Typha latifolia to evaluate its bioenergy potential. Bioresour. Technol. 2017, 245, 491–501. [Google Scholar] [CrossRef] [Green Version]
  21. Kumar, M.; Sabbarwal, S.; Mishra, P.; Upadhyay, S. Thermal degradation kinetics of sugarcane leaves (Saccharum officinarum L) using thermo-gravimetric and differential scanning calorimetric studies. Bioresour. Technol. 2019, 279, 262–270. [Google Scholar] [CrossRef]
  22. Chandrasekaran, A.; Ramachandran, S.; Subbiah, S. Determination of kinetic parameters in the pyrolysis operation and thermal behavior of Prosopis juliflora using thermogravimetric analysis. Bioresour. Technol. 2017, 233, 413–422. [Google Scholar] [CrossRef]
  23. Mishra, R.K.; Mohanty, K. Kinetic analysis and pyrolysis behaviour of waste biomass towards its bioenergy potential. Bioresour. Technol. 2020, 311, 123480. [Google Scholar] [CrossRef] [PubMed]
  24. Soria-Verdugo, A.; Goos, E.; García-Hernando, N.; Riedel, U. Analyzing the pyrolysis kinetics of several microalgae species by various differential and integral isoconversional kinetic methods and the Distributed Activation Energy Model. Algal Res. 2018, 32, 11–29. [Google Scholar] [CrossRef] [Green Version]
  25. Ermolaev, D.V.; Timofeeva, S.S.; Islamova, S.I.; Bulygina, K.S.; Gilfanov, M.F. A comprehensive study of thermotechnical and thermogravimetric properties of peat for power generation. Biomass Conv. Bioref. 2019, 9, 767–774. [Google Scholar] [CrossRef]
  26. Karaeva, J.V.; Timofeeva, S.S.; Bashkirov, V.N.; Bulygina, K.S. Thermochemical processing of digestate from biogas plant for recycling dairy manure and biomass. Biomass Conv. Bioref. 2021, 13, 685–695. [Google Scholar] [CrossRef]
  27. Wang, G.; Dai, Y.; Yang, H.; Xiong, Q.; Wang, K.; Zhou, J.; Li, Y.; Wang, S. A Review of Recent Advances in Biomass Pyrolysis. Energy Fuels 2020, 34, 15557–15578. [Google Scholar] [CrossRef]
  28. Lewandowski, W.M.; Ryms, M.; Kosakowski, W. Thermal Biomass Conversion: A Review. Processes 2020, 8, 516. [Google Scholar] [CrossRef]
  29. Karaeva, J.V.; Timofeeva, S.S.; Kovalev, A.A.; Kovalev, D.A.; Gilfanov, M.F.; Grigoriev, V.S.; Litti, Y.V. Co-pyrolysis of agri-cultural waste and estimation of the applicability of pyrolysis in the integrated technology of biorenewable hydrogen produc-tion. Int. J. Hydrogen Energy 2022, 47, 11787–11798. [Google Scholar] [CrossRef]
  30. Sun, Y.; Liu, K.; Kun, L.; Hou, C.; Liu, J.; Huang, R.; Cao, C.; Song, W. Nitrogen, Sulfur Co-doped Carbon Materials Derived from the Leaf, Stem and Root of Amaranth as Metal-free Catalysts for Selective Oxidation of Aromatic Hydrocarbons. ChemCatChem 2019, 11, 1010–1016. [Google Scholar] [CrossRef]
  31. Gao, S.; Geng, K.; Liu, H.; Wei, X.; Zhang, M.; Wang, P.; Wang, J. Transforming organic-rich amaranthus waste into nitro-gen-doped carbon with superior performance of the oxygen reduction reaction. Energy Environ. Sci. 2015, 8, 221–229. [Google Scholar] [CrossRef]
  32. Tinwala, F.; Mohanty, P.; Parmar, S.; Patel, A.; Pant, K.K. Intermediate pyrolysis of agro-industrial biomasses in bench-scale pyrolyser: Product yields and its characterization. Bioresour. Technol. 2015, 188, 258–264. [Google Scholar] [CrossRef]
  33. Biswas, B.; Pandey, N.; Bisht, Y.; Singh, R.; Kumar, J.; Bhaskar, T. Pyrolysis of agricultural biomass residues: Comparative study of corn cob, wheat straw, rice straw and rice husk. Bioresour. Technol. 2017, 237, 57–63. [Google Scholar] [CrossRef] [PubMed]
  34. Rego, F.; Dias, A.P.S.; Casquilho, M.; Rosa, F.C.; Rodrigues, A. Pyrolysis kinetics of short rotation coppice poplar biomass. Energy 2020, 207, 118191. [Google Scholar] [CrossRef]
  35. Bhardwaj, G.; Kumar, M.; Mishra, P.K.; Upadhyay, S.N. Kinetic analysis of the slow pyrolysis of paper wastes. Biomass Conv. Bioref. 2021, 11, 1–14. [Google Scholar] [CrossRef]
  36. Shahbeig, H.; Nosrati, M. Pyrolysis of biological wastes for bioenergy production: Thermo-kinetic studies with machine-learning method and Py-GC/MS analysis. Fuel 2020, 269, 117238. [Google Scholar] [CrossRef]
  37. Singh, R.K.; Pandey, D.; Patil, T.; Sawarkar, A.N. Pyrolysis of banana leaves biomass: Physico-chemical characterization, thermal decomposition behavior, kinetic and thermodynamic analyses. Bioresour. Technol. 2020, 310, 123464. [Google Scholar] [CrossRef]
  38. Jiang, G.; Wei, L. Analysis of pyrolysis kinetic model for processing of thermogravimetric analysis data. In Phase Change Materials and Their Applications; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef] [Green Version]
  39. Yuan, X.; He, T.; Cao, H.; Yuan, Q. Cattle manure pyrolysis process: Kinetic and thermodynamic analysis with isoconversional methods. Renew. Energy 2017, 107, 489–496. [Google Scholar] [CrossRef]
  40. Khawam, A. Application of Solid-State Kinetics to Desolvation Reactions. Ph.D. Thesis, University of Iowa, Iowa, IA, USA, 2007. [Google Scholar]
  41. Sánchez-Jiménez, P.E.; Pérez-Maqueda, L.A.; Perejón, A.; Criado, J.M. Generalized master plots as a straightforward approach for determining the kinetic model: The case of cellulose pyrolysis. Thermochim. Acta 2013, 552, 54–59. [Google Scholar] [CrossRef] [Green Version]
  42. Gotor, F.J.; Criado, J.M.; Malek, J.; Koga, N. Kinetic Analysis of Solid-State Reactions: The Universality of Master Plots for Analyzing Isothermal and Nonisothermal Experiments. J. Phys. Chem. A 2000, 104, 10777–10782. [Google Scholar] [CrossRef]
  43. Brown, M.E. Handbook of Thermal Analysis and Calorimetry; Principles and Practice; Elsevier Science: Amsterdam, The Netherlands, 1998; Volume 1, p. 722. [Google Scholar]
  44. Alves, J.L.F.; Da Silva, J.C.G.; Filho, V.F.S.; Alves, R.F.; Ahmad, M.S.; Ahmad, M.S.; Galdino, W.V.A.; De Sena, R.F. Bioenergy potential of red macroalgaeGelidium floridanumby pyrolysis: Evaluation of kinetic triplet and thermodynamics parameters. Bioresour. Technol. 2019, 291, 121892. [Google Scholar] [CrossRef]
  45. Varma, A.K.; Lal, N.; Rathore, A.K.; Katiyar, R.; Thakur, L.S.; Shankar, R.; Mondal, P. Thermal, kinetic and thermodynamic study for co-pyrolysis of pine needles and styrofoam using thermogravimetric analysis. Energy 2020, 218, 119404. [Google Scholar] [CrossRef]
  46. El-Sayed, S.A.; Mostafa, M.E. Kinetics, thermodynamics, and combustion characteristics of Poinciana pods using TG/DTG/DTA techniques. Biomass Conv. Bioref. 2021, 11, 1–25. [Google Scholar] [CrossRef]
  47. García, R.; Pizarro, C.; Lavín, A.G.; Bueno, J.L. Characterization of Spanish biomass wastes for energy use. Bioresour. Technol. 2012, 103, 249–258. [Google Scholar] [CrossRef] [PubMed]
  48. Emiola-Sadiq, T.; Zhang, L.; Dalai, A.K. Thermal and Kinetic Studies on Biomass Degradation via Thermogravimetric Anal-ysis: A Combination of Model-Fitting and Model-Free Approach. ACS Omega 2021, 6, 22233–22247. [Google Scholar] [CrossRef] [PubMed]
  49. Güleç, F.; Pekaslan, D.; Williams, O.; Lester, E. Predictability of higher heating value of biomass feedstocks via proximate and ultimate analyses—A comprehensive study of artificial neural network applications. Fuel 2022, 320, 123944. [Google Scholar] [CrossRef]
  50. Imam, T.; Capareda, S. Characterization of bio-oil, syn-gas and bio-char from switchgrass pyrolysis at various temperatures. J. Anal. Appl. Pyrolysis 2012, 93, 170–177. [Google Scholar] [CrossRef]
  51. Aboulkas, A.; Hammani, H.; El Achaby, M.; Bilal, E.; Barakat, A.; El harfi, K. Valorization of algal waste via pyrolysis in a fixed-bed reactor: Production and characterization of bio-oil and bio-char. Bioresour. Technol. 2017, 243, 400–408. [Google Scholar] [CrossRef]
  52. Pandey, D.S.; Katsaros, G.; Lindfors, C.; Leahy, J.J.; Tassou, S.A. Fast Pyrolysis of Poultry Litter in a Bubbling Fluidised Bed Reactor: Energy and Nutrient Recovery. Sustainability 2019, 11, 2533. [Google Scholar] [CrossRef] [Green Version]
  53. Panchasara, H.; Ashwath, N. Effects of Pyrolysis Bio-Oils on Fuel Atomisation—A Review. Energies 2021, 14, 794. [Google Scholar] [CrossRef]
  54. Alvarez, J.; Lopez, G.; Amutio, M.; Artetxe, M.; Barbarias, I.; Arregi, A.; Bilbao, J.; Olazar, M. Characterization of the bio-oil obtained by fast pyrolysis of sewage sludge in a conical spouted bed reactor. Fuel Process. Technol. 2016, 149, 169–175. [Google Scholar] [CrossRef]
  55. Nasirpour-Tabrizi, P.; Azadmard-Damirchi, S.; Hesari, J.; Piravi-Vanak, Z. Amaranth Seed Oil Composition; Waisundara, V.Y., Ed.; IntechOpen: London, UK, 2020. [Google Scholar]
  56. Griessacher, T.; Antrekowitsch, J.; Steinlechner, S. Charcoal from agricultural residues as alternative reducing agent in metal recycling. Biomass Bioenergy 2012, 39, 139–146. [Google Scholar] [CrossRef]
  57. Peer, V.; Frantík, J.; Kielar, J.; Mašek, D. Substrates for slow pyrolysis. In Proceedings of the XXI International Scientific Conference—The Application of Experimental and Numerical Methods in Fluid Mechanics and Energy, Teplice, Slovakia, 25–27 April 2018. [Google Scholar]
  58. Das, S.K.; Ghosh, G.K.; Avasthe, R.K.; Sinha, K. Compositional heterogeneity of different biochar: Effect of pyrolysis temper-ature and feedstocks. J. Environ. Manag. 2021, 278, 2. [Google Scholar] [CrossRef] [PubMed]
  59. Tangmankongworakoon, N. An approach to produce biochar from coffee residue for fuel and soil amendment purpose. Int. J. Recycl. Org. Waste Agric. 2019, 8 (Suppl. S1), 37–44. [Google Scholar] [CrossRef] [Green Version]
  60. Baidoo, I.K.; Sarpong, D.B.; Bolwig, S.; Ninson, D. Biochar amended soils and crop productivity: A critical and meta-analysis of literature. Int. J. Dev. Sustain. 2016, 5, 414–432. [Google Scholar]
  61. Yuan, J.-H.; Xu, R.-K.; Qian, W.; Wang, R.-H. Comparison of the ameliorating effects on an acidic ultisol between four crop straws and their biochars. J. Soils Sediments 2011, 11, 741–750. [Google Scholar] [CrossRef]
  62. Anca-Couce, A. Reaction mechanisms and multi-scale modelling of lignocellulosic biomass pyrolysis. Prog. Energy Combust. Sci. 2016, 53, 41–79. [Google Scholar] [CrossRef]
  63. Pielsticker, S.; Gövert, B.; Umeki, K.; Kneer, R. Flash Pyrolysis Kinetics of Extracted Lignocellulosic Biomass Components. Front. Energy Res. 2021, 9, 497. [Google Scholar] [CrossRef]
  64. Senneca, O.; Cerciello, F.; Russo, C.; Wütscher, A.; Muhler, M.; Apicella, B. Thermal treatment of lignin, cellulose and hemi-cellulose in nitrogen and carbon dioxide. Fuel 2020, 271, 117656. [Google Scholar] [CrossRef]
  65. Basu, P. Biomass Gasification, Pyrolysis and Torrefaction, 3rd ed.; Academic Press: Cambridge, MA, USA, 2018; p. 564. [Google Scholar]
  66. Shen, D.K.; Gu, S.; Bridgwater, A.V. The thermal performance of the polysaccharides extracted from hardwood: Cellulose and hemicellulose. Carbohydr. Polym. 2010, 82, 39–45. [Google Scholar] [CrossRef]
  67. Usino, D.O.; Ylitervo, P.; Moreno, A.; Sipponen, M.H.; Richards, T. Primary interactions of biomass components during fast pyrolysis. J. Anal. Appl. Pyrolysis 2021, 159, 105297. [Google Scholar] [CrossRef]
  68. Yu, J.; Paterson, N.; Blamey, J.; Millan, M. Cellulose, xylan and lignin interactions during pyrolysis of lignocellulosic biomass. Fuel 2017, 191, 140–149. [Google Scholar] [CrossRef]
  69. Shen, D.; Xiao, R.; Gu, S.; Luo, K. The pyrolytic behavior of cellulose in lignocellulosic biomass: A review. RSC Adv. 2011, 1, 1641–1660. [Google Scholar] [CrossRef]
  70. Qiao, Y.; Wang, B.; Ji, Y.; Xu, F.; Zong, P.; Zhang, J.; Tian, Y. Thermal decomposition of castor oil, corn starch, soy protein, lignin, xylan, and cellulose during fast pyrolysis. Bioresour. Technol. 2019, 278, 287–295. [Google Scholar] [CrossRef] [PubMed]
  71. Wang, S.; Lin, H.; Ru, B.; Sun, W.; Wang, Y.; Luo, Z. Comparison of the pyrolysis behavior of pyrolytic lignin and milled wood lignin by using TG–FTIR analysis. J. Anal. Appl. Pyrolysis 2014, 108, 78–85. [Google Scholar] [CrossRef]
  72. Peng, C.; Zhang, G.; Yue, J.; Xu, G. Pyrolysis of black liquor for phenols and impact of its inherent alkali. Fuel Process. Technol. 2014, 127, 149–156. [Google Scholar] [CrossRef]
  73. Lin, F.; Waters, C.L.; Mallinson, R.G.; Lobban, L.L.; Bartley, L.E. Relationships between Biomass Composition and Liquid Products Formed via Pyrolysis. Front. Energy Res. 2015, 3, 45. [Google Scholar] [CrossRef] [Green Version]
  74. Laougé, Z.B.; Merdun, H. Pyrolysis and combustion kinetics of Sida cordifolia L. using thermogravimetric analysis. Bioresour. Technol. 2020, 299, 122602. [Google Scholar] [CrossRef]
  75. Vuppaladadiyam, A.K.; Antunes, E.; Sanchez, P.B.; Duan, H.; Zhao, M. Influence of microalgae on synergism during co-pyrolysis with organic waste biomass: A thermogravimetric and kinetic analysis. Renew. Energy 2020, 167, 42–55. [Google Scholar] [CrossRef]
  76. Mallick, D.; Bora, B.J.; Baruah, D.; Barbhuiya, S.A.; Banik, R.; Garg, J.; Sarma, R. Mechanistic investigation and thermal deg-radation of Eichhornia crassipes using Thermogravimetric analysis. SSRN Electron. J. 2020. [Google Scholar] [CrossRef]
  77. Baruah, D.; Mallick, D.; Kalita, P.; Moholkar, V.S. A Detailed Study of Pyrolysis Kinetics of Elephant Grass Using Thermogravimetric Analysis. In Proceedings of the 2nd International Conference on Energy Power and Environment, Shillong, India, 1–2 June 2018. [Google Scholar]
  78. Patidar, K.; Singathia, A.; Vashishtha, M.; Sangal, V.K.; Upadhyaya, S. Investigation of kinetic and thermodynamic parameters approaches to non-isothermal pyrolysis of mustard stalk using model-free and master plots methods. Mater. Sci. Energy Technol. 2021, 5, 6–14. [Google Scholar] [CrossRef]
  79. Fernandez, A.; Ortiz, L.R.; Asensio, D.; Rodriguez, R.; Mazza, G. Kinetic analysis and thermodynamics properties of air/steam gasification of agricultural waste. J. Environ. Chem. Eng. 2020, 8, 103829. [Google Scholar] [CrossRef]
  80. Wang, S.; Wen, Y.; Shi, Z.; Niedzwiecki, L.; Baranowski, M.; Czerep, M.; Mu, W.; Kruczek, H.P.; Jönsson, P.G.; Yang, W. Effect of hydrothermal carbonization pretreatment on the pyrolysis behavior of the digestate of agricultural waste: A view on kinetics and thermodynamics. Chem. Eng. J. 2022, 431, 133881. [Google Scholar] [CrossRef]
  81. Ozawa, T. Estimation of activation energy by isoconversion methods. Thermochim. Acta 1992, 203, 159–165. [Google Scholar] [CrossRef]
  82. Mandapati, R.N.; Ghodke, P.K. Kinetics of pyrolysis of cotton stalk using model-fitting and model-free methods. Fuel 2021, 303, 121285. [Google Scholar] [CrossRef]
  83. 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]
  84. Hilten, R.; Vandenbrink, J.; Paterson, A.; Feltus, F.; Das, K. Linking isoconversional pyrolysis kinetics to compositional characteristics for multiple Sorghum bicolor genotypes. Thermochim. Acta 2013, 577, 46–52. [Google Scholar] [CrossRef]
  85. Müsellim, E.; Tahir, M.H.; Ahmad, M.S.; Ceylan, S. Thermokinetic and TG/DSC-FTIR study of pea waste biomass pyrolysis. Appl. Therm. Eng. 2018, 137, 54–61. [Google Scholar] [CrossRef]
  86. Sharma, A.; Mohantya, B. Thermal degradation of mango (Mangifera indica) wood sawdust in a nitrogen environment: Characterization, kinetics, reaction mechanism, and thermodynamic analysis. RSC Adv. 2021, 11, 13396–13408. [Google Scholar] [CrossRef]
  87. Phuakpunk, K.; Chalermsinsuwan, B.; Assabumrungrat, S. Pyrolysis kinetic parameters investigation of single and tri-component biomass: Models fitting via comparative model-free methods. Renew. Energy 2022, 182, 494–507. [Google Scholar] [CrossRef]
  88. Colpani, D.; Santos, V.O.; Araujo, R.O.; Lima, V.M.R.; Tenório, J.A.S.; Coleti, J.; Chaar, J.S.; de Souza, L.K.C. Bioenergy po-tential analysis of Brazil nut biomass residues through pyrolysis: Gas emission, kinetics, and thermodynamic parameters. Clean. Chem. Eng. 2022, 1, 100002. [Google Scholar]
  89. Poletto, M.; Zattera, A.J.; Santana, R.M. Thermal decomposition of wood: Kinetics and degradation mechanisms. Bioresour. Technol. 2012, 126, 7–12. [Google Scholar] [CrossRef]
  90. Mallick, D.; Poddar, M.K.; Mahanta, P.; Moholkar, V.S. Discernment of synergism in pyrolysis of biomass blends using ther-mogravimetric analysis. Bioresour. Technol. 2018, 261, 294–305. [Google Scholar] [CrossRef] [PubMed]
  91. Vikraman, V.K.; Boopathi, G.; Kumar, D.P.; Mythili, R.; Subramanian, P. Non-isothermal pyrolytic kinetics of milk dust powder using thermogravimetric analysis. Renew. Energy 2021, 180, 838–849. [Google Scholar] [CrossRef]
  92. Kumar, M.; Mishra, P.K.; Upadhyay, S.N. Thermal degradation of rice husk: Effect of pre-treatment on kinetic and thermo-dynamic parameters. Fuel 2020, 268, 117164. [Google Scholar] [CrossRef]
  93. Postawa, K.; Fałtynowicz, H.; Szczygieł, J.; Beran, E.; Kułażyński, 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]
  94. Xiao, R.; Yang, W.; Cong, X.; Dong, K.; Xu, J.; Wang, D.; Yang, X. Thermogravimetric analysis and reaction kinetics of ligno-cellulosic biomass pyrolysis. Energy 2020, 201, 117537. [Google Scholar] [CrossRef]
  95. Maia, A.A.D.; De Morais, L.C. Kinetic parameters of red pepper waste as biomass to solid biofuel. Bioresour. Technol. 2016, 204, 157–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Sahoo, A.; Kumar, S.; Mohanty, K. Kinetic and thermodynamic analysis of Putranjiva roxburghii (putranjiva) and Cassia fistula (amaltas) non-edible oilseeds using thermogravimetric analyzer. Renew. Energy 2021, 165, 261–277. [Google Scholar] [CrossRef]
  97. Hu, L.; Wei, X.-Y.; Guo, X.-H.; Lv, H.-P.; Wang, G.-H. Investigation on the kinetic behavior, thermodynamic and volatile products analysis of chili straw waste pyrolysis. J. Environ. Chem. Eng. 2021, 9, 105859. [Google Scholar] [CrossRef]
  98. Mishra, A.; Kumari, U.; Turlapati, V.Y.; Siddiqi, H.; Meikap, B. Extensive thermogravimetric and thermo-kinetic study of waste motor oil based on iso-conversional methods. Energy Convers. Manag. 2020, 221, 113194. [Google Scholar] [CrossRef]
  99. Simões, L.M.S.; Setter, C.; Sousa, N.G.; Cardoso, C.R.; de Oliveira, T.J.P. Biomass to biofuel densification of coconut fibers: Kinetic triplet and thermodynamic evaluation. Biomass Conv. Bioref. 2022, 12. [Google Scholar] [CrossRef]
  100. Singh, S.; Chakraborty, J.P.; Mondal, M.K. Intrinsic kinetics, thermodynamic parameters and reaction mechanism of non-isothermal degradation of torrefied Acacia nilotica using isoconversional methods. Fuel 2020, 259, 116263. [Google Scholar] [CrossRef]
  101. Muigai, H.H.; Choudhury, B.J.; Kalita, P.; Moholkar, V.S. Co–pyrolysis of biomass blends: Characterization, kinetic and thermodynamic analysis. Biomass Bioenergy 2020, 143, 105839. [Google Scholar] [CrossRef]
Figure 1. AIW pyrolysis products.
Figure 1. AIW pyrolysis products.
Agriculture 13 00260 g001
Figure 2. Pyrolysis Liquid: (a) photograph and (b) composition of the oil fraction.
Figure 2. Pyrolysis Liquid: (a) photograph and (b) composition of the oil fraction.
Agriculture 13 00260 g002
Figure 3. Biochar: (a) photography; (b) elemental composition of mineral part.
Figure 3. Biochar: (a) photography; (b) elemental composition of mineral part.
Agriculture 13 00260 g003
Figure 4. (a) TG−curves; (b) DTG−curves.
Figure 4. (a) TG−curves; (b) DTG−curves.
Agriculture 13 00260 g004
Figure 5. Plots for determination of AIW pyrolysis using: (a) Friedman; (b) KAS; (c) OFW.
Figure 5. Plots for determination of AIW pyrolysis using: (a) Friedman; (b) KAS; (c) OFW.
Agriculture 13 00260 g005
Figure 6. Comparison of experimental and theoretical master plots for samples AIW.
Figure 6. Comparison of experimental and theoretical master plots for samples AIW.
Agriculture 13 00260 g006
Figure 7. (a) ∆H, (b) ∆G, and (c) ∆S of the pyrolysis of AIW.
Figure 7. (a) ∆H, (b) ∆G, and (c) ∆S of the pyrolysis of AIW.
Agriculture 13 00260 g007aAgriculture 13 00260 g007b
Table 1. The results of the proximate and ultimate analyses of AIW sample.
Table 1. The results of the proximate and ultimate analyses of AIW sample.
AnalysisValues
Proximate (wt.%)—based on air-dried basis:
Moisture7.42 ± 0.02
Volatile matter74.65 ± 0.30
Ash8.76 ± 0.01
Fixed carbon9.17 ± 0.06
HHV, MJ/kg17.87
Ultimate (wt.%)—based on dry basis:
Carbon41.83 ± 0.26
Hydrogen6.81 ± 0.08
Nitrogen4.71 ± 0.13
Oxygen37.89 ± 0.17
Table 2. The main components of the oil fraction (peak area ≥ 1%).
Table 2. The main components of the oil fraction (peak area ≥ 1%).
Area, %NameFormulaMw, g/mol
112.36TetratetracontaneC44H90619.8
29.70TetracontaneC40H82563.1
38.201-OctacosanolC28H58O410.8
45.442,6,10,14,18,22-Tetracosahexaene, 2,6,10,15,19,23-hexamethyl-, (all-E)-C30H50410.7
52.19Octacosanoic acid, methyl esterC29H58O2438.8
62.00PhenolC6H5OH94.11
71.80PentadecaneC15H32212.41
81.50Triacontanoic acid, methyl esterC31H62O2466.82
91.43TetracosaneC24H50338.7
101.40Phenol, 2-methoxy-C7H8O2124.12
111.24Pyridine, 3-methyl-C6H7N93.13
121.11OctadecaneC18H38254.49
Table 3. Pyro—gas composition.
Table 3. Pyro—gas composition.
ComponentCO2COCH4C2H6CxHyC2H4H2
Concentration, %47.3047.143.511.170.750.120.01
Table 4. The results of the proximate and ultimate analyses for biomass biochars.
Table 4. The results of the proximate and ultimate analyses for biomass biochars.
AnalysisBiomass
AIWMaize Stalk
[58]
Lantana Camara
[58]
Pine Needles [58]Black Gram [58]
Proximate (wt.%)
Volatile matter21.34 ± 0.0320.6722.5627.6223.56
Ash20.49 ± 0.0119.715.713.523.3
Moisture4.4 ± 0.1911.56.138.0512.41
Fixed carbon53.77 ± 0.0,948.1355.6150.8340.73
HHV *, MJ/kg20.9223.725.8722.3321.06
Ultimate (wt.%)
carbon56.56 ± 0.1761.970.565.856.7
hydrogen3.09 ± 0.053.562.692.133.14
nitrogen4.12 ± 0.011.170.860.781.24
oxygen15.75 ± 0.1413.6710.2517.7915.62
* calculated.
Table 5. Main stages of thermal decomposition.
Table 5. Main stages of thermal decomposition.
Pyrolysis StageHeating Rate (°C/min)Starting Temperature (°C)Ending
Temperature (°C)
Temperature Peak (°C)
IMoisture evaporation10
15
20
40
40
40
191.26
190.76
191.77
103.1
115.4
126.7
IIDevolatilization10
15
20
191.26
190.76
191.77
529.5
544.48
558.95
317.7
322.6
328.5
IIIDegradation of char and minerals10
15
20
529.5
544.48
558.95
1000
1000
1000
668.2
685.6
690.6
Table 6. Mass loss characteristics of AIW obtained from TGA analysis.
Table 6. Mass loss characteristics of AIW obtained from TGA analysis.
Heating Rate (°C/min)Mass Loss, wt.%Residual Mass, wt.%
Moisture EvaporationDevolatilizationDegradation of Char and Minerals
109.1458.655.7926.24
159.3359.175.8925.61
209.2861.084.8824.76
Average, %9.2559.635.5225.54
Table 7. Values of activated energy according to different methods.
Table 7. Values of activated energy according to different methods.
αFriedmanKASOFW
Eα (kJ/mol)Log A (1/s)Eα (kJ/mol)Log A (1/s)Eα (kJ/mol)Log A (1/s)
0.1164.4813.22185.2715.72185.4215.68
0.2152.5211.44156.7812.26157.0012.23
0.3184.7714.16167.3112.93167.5012.90
0.4212.8816.47189.5714.76189.7414.72
0.5231.0317.80209.0816.31209.2316.26
0.6250.6019.10225.7617.52225.9117.48
0.7291.9421.85253.7319.48253.8819.44
0.8244.2916.61265.2519.30265.4319.26
0.9213.0213.17223.2114.75223.4314.70
Average216.1715.98208.4415.89208.6115.85
Table 8. Comparison of biomasses activation energy.
Table 8. Comparison of biomasses activation energy.
FuelHeating Rate (K/min)Used MethodsActivation Energy (kJ/mol)Reference
AIW10, 15, and 20Friedman,
KAS,
OFW
216.17
208.44
208.61
Present Study
Cotton stalk10–40KAS,
OFW
223–230
213–240
[82]
Sugarcane leaves5–40Friedman,
KAS,
OFW
239.58
226.75
226.97
[21]
Prosopis juliflora fuelwood2–25Friedman,
KAS,
OFW
219.3
204.0
203.2
[22]
Phyllanthus emblica seeds10–50Friedman,
KAS,
OFW
189.95
184.77
195.10
[23]
Camphor branch2.5, 5, and 10Ozawa190[83]
Microalgae Chlorella vulgaris10–40Kissinger, Friedman, OFW, KAS, Vyazovkin, DAEM135.6–337.1[24]
Digested biomass wastes10, 15, and 20Friedman,
KAS
202.55
202.21
[75]
Sorghum bicolor2, 5, and 8Friedman and KAS226.6[84]
Pea waste10–40KAS,
OFW
212.71
211.55
[85]
Basswood waste20–40KAS,
OFW
197.2
207.9
[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

Karaeva, J.; Timofeeva, S.; Islamova, S.; Bulygina, K.; Aliev, F.; Panchenko, V.; Bolshev, V. Pyrolysis of Amaranth Inflorescence Wastes: Bioenergy Potential, Biochar and Hydrocarbon Rich Bio-Oil Production. Agriculture 2023, 13, 260. https://doi.org/10.3390/agriculture13020260

AMA Style

Karaeva J, Timofeeva S, Islamova S, Bulygina K, Aliev F, Panchenko V, Bolshev V. Pyrolysis of Amaranth Inflorescence Wastes: Bioenergy Potential, Biochar and Hydrocarbon Rich Bio-Oil Production. Agriculture. 2023; 13(2):260. https://doi.org/10.3390/agriculture13020260

Chicago/Turabian Style

Karaeva, Julia, Svetlana Timofeeva, Svetlana Islamova, Kseny Bulygina, Firdavs Aliev, Vladimir Panchenko, and Vadim Bolshev. 2023. "Pyrolysis of Amaranth Inflorescence Wastes: Bioenergy Potential, Biochar and Hydrocarbon Rich Bio-Oil Production" Agriculture 13, no. 2: 260. https://doi.org/10.3390/agriculture13020260

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