Next Article in Journal
Investigation on Deformation Behavior of the Crossing Section of Two Municipal Road Tunnels during Construction
Previous Article in Journal
Suitability of an Artificial Viscosity Model for Compressible Under-Resolved Turbulence Using a Flux Reconstruction Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Mathematical Model of Industrial Waste-Derived Fuel Droplet Combustion in High-Temperature Air

by
Dmitrii Antonov
1,
Dmitrii Glushkov
1,*,
Kristina Paushkina
1,
Daria Kuznechenkova
1 and
Anand Ramanathan
2
1
Heat Mass Transfer Laboratory, National Research Tomsk Polytechnic University, Tomsk 634050, Russia
2
Department of Mechanical Engineering, National Institute of Technology Tiruchirappalli, Tiruchirappalli 620015, India
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(23), 12273; https://doi.org/10.3390/app122312273
Submission received: 27 October 2022 / Revised: 23 November 2022 / Accepted: 28 November 2022 / Published: 30 November 2022
(This article belongs to the Section Energy Science and Technology)

Abstract

:
The results of a theoretical and experimental study of the processes of ignition and combustion of a single composite liquid fuel (CLF) droplet based on wet coal processing waste and combustible municipal solid waste under radiant and convective heating are presented. Based on the results of a detailed analysis of video recordings and previously obtained experimental data from the ignition and combustion of a single CLF droplet, a mathematical model was developed. The advantage of the developed mathematical model lies in the specification of sequential physical and chemical processes of the high-temperature decomposition of fuel in a high-temperature gaseous medium. A numerical simulation of combustion characteristics was carried out in the Ansys Fluent commercial software for five different CLF compositions. The ignition-delay times were established for fuel droplets that were in a preheated motionless air environment, a temperature in the range of 723–1273 K, and an air flow heated to 723–973 K moving at a velocity of 3 m/s. Using the asymptotic procedure, satisfactory analytical solutions are obtained for the multistage nonlinear problem of ignition and combustion of a single CLF droplet. The possibility for the practical application of the developed program in Ansys Fluent in predicting the characteristics of the ignition processes of CLF droplets is substantiated.

1. Introduction

The current hazards of fossil fuel depletion in the medium term and of irreversible climate change due to global warming have prompted the design and implementation of sustainable development policy in the main areas of the economy. Quickly growing population, rapid industrialization and global economic changes have increased the demand for electric and thermal energy [1] even as there is a projected deficit of energy resources and fossil fuels. According to the industry estimates [2], even though there was a decrease in the energy demands during the COVID-19 pandemic, humanity is expected to consume 10% more primary energy in 2030 (and 35% more in 2050) compared to the annual consumption in 2019.
Coal energy production is a globally widespread technology that generated about seven billion tons of industrial waste in 2020 alone [3], and this figure is projected to rise to 25 billion tons by 2026 [4]. In addition to industrial waste, municipal waste (including combustible waste) pollutes the environment. According to estimates [5], if humanity annually produced 2.1 billion tons of municipal solid waste (MSW) in 2019, then by 2050 people will generate about 3.4 billion tons. Stored at open-air disposal sites, industrial and municipal solid wastes pollute the environment (air, soil and water) as a result of their slow natural decay [5,6]. Overall, this leads to deteriorating conditions for both people and wildlife living in the vicinity [7,8,9,10].
One of the most economically sparing ways to utilize accumulated and unclaimed waste is the method of burning composite liquid fuels (CLF). The results of a cycle of experimental research [11,12,13] into the combustion of CLF, consisting of wet coal enrichment waste (high-moisture low-rank coal) and fine MSW, have proved the efficiency of using the waste-to-energy technology to generate heat and power with a relatively low level of anthropogenic emissions.
The use of such CLFs in practice is limited by the need to develop a number of compositions with predictable energy characteristics. The scientific justification of a high-potential fuel component composition is time-consuming, costly and challenging. The numerical simulation of physical and chemical processes is an alternative approach to the efficient prediction of fuel combustion characteristics under different conditions.
In recent years, many mathematical models have been developed to study the influence of environmental conditions and fuel properties on the thermal decomposition characteristics of different grades of coal [14,15] and a group of typical MSW components taken separately [16,17,18,19]. Studies on the modeling of MSW thermal decomposition for subsequent practical application generally state the problem with significant assumptions and simplifications, which increases the scale of the problem to be solved and lowers the requirements of the numerical simulation algorithm for the PC’s computational resources. Yet this approach does not make it possible to explore the physics and chemistry of solid fuel conversion. Studies like that may use empirical models of the conversion of solid fuel in a densified layer divided into several zones (e.g., drying, pyrolysis, combustion, gasification), to estimate the qualitative and quantitative characteristics of gaseous pyrolysis and gasification products [18].
In a number of works [17,19], the Eulerian–Lagrangian model is used. It describes gas (liquid) flows with a low content of the dispersed phase [20,21]. Solid fuel particles are considered as a discrete phase (Lagrangian), and gas (liquid) flows are considered as a continuous medium (Eulerian). However, the Eulerian–Lagrangian approach is not suitable for modeling the processes in suspensions, fluidized bed layers and any other mixtures in which the volume ratio of the solid phase cannot be neglected. Fuel conversion in the fluidized bed is described using the Eulerian–Eulerian model (continuum model) [16,22], which describes different phases as continuous media. Such a simplified approach makes it possible to reliably predict the temperature fields and velocity profiles, though it is not relevant for an in-depth study of heat and mass transfer, as well as hydrodynamics processes developing in a single fuel particle and in its vicinity.
Analysis of the synergistic effect of the components of the CLF mixture (coal, water and MSW) on the temporal and environmental characteristics of the fuel combustion is a promising direction with significant potential for the development of previously obtained experimental research findings [23]. Therefore, the aim of this work is to develop, on the basis of the results of previous experimental studies [23,24], a mathematical model for the combustion of a unit multicomponent CLF droplet in an air medium heated to high temperatures, which provides a satisfactory reproduction of the experimental data. The model is supposed to intensify research in the field of eco-energy and to develop a strategy for the sustainability of the electric power industry with limited consumption of exhaustible natural resources and minimal negative impact on the environment.

2. Materials and Methods

2.1. Background of the Model Condition

The ability of CLS to reliably ignite and burn out completely in a preheated gaseous medium has been established by us experimentally [23]. The identified mechanism and recorded characteristics of the process set the stage for the development of the mathematical model and numerical simulation of single fuel droplet combustion in Ansys Fluent. The main component of the CLF was filter cake (FC) of coking coal C from the Severnaya coal plant (Kuzbass, Russian Federation) [23]. FC is a waste from the coal industry, obtained in the process of coal enrichment. FC has an energy potential due to the content of at least 50% fine (about 100 µm) coal in the composition (the rest is water) [25,26,27]. However, so far, FC is not widely used in any industries, so the main method of its disposal is uncontrolled storage on open sites near the washing plant. This method of coal waste disposal is characterized by many negative consequences, primarily environmental. Moisture evaporation increases the fire hazard of coal dust, allows it to be dispersed by the wind over large areas, polluting the air, open water bodies and soils. Moreover, coal dust contains acidifying agents, high-density metals and other hazardous elements. These become a source of integrated pollution for the environment, when exposed to a severe impact of natural factors (such as air, moisture or solar energy).
Besides FC, combustible MSW can also be a CLF component [23,28]. MSW typically consists [5,29,30] of 26–35% organic remains (including former foodstuff); 25–30% (of the total waste volume) mixed paper wastage; 7–10% postconsumer plastics; 5–12% scrap metal and cullet; 2–4% textile wastes, 2–4% wood residue, and 2–4% all-rubber scraps. Based on statistical data, MSW consists of about 82% potentially energy-producing materials (paper, wood, textile, polymer waste). Adding the combustible components of MSW to composite fuel is a promising solution to the problem of commercial disposal of accumulated and newly produced waste. However, the relatively low energy content of combustible MSW (10 MJ/kg [31,32,33] versus 20–30 MJ/kg [34] of solid fossil fuels) and a fairly high concentration of hazardous compounds (including carcinogens) limit the amount that can be added to a CLF composition to 10–20% [23].
The main properties, characteristics and ultimate analysis of CLF are presented in Table 1 and Table 2 [28]. To obtain the characteristics and properties of the CLF components, the samples were preliminarily dried at a temperature of about 378 K to evaporate unbound moisture.
In the experimental research [23], MSW characteristics from reference materials were used to analyze the effect of CLF component properties on the patterns of physical and chemical processes with different mechanisms of heat supply [31,35,36]. MSW composition and characteristics may vary considerably in different regions (by methods of waste collection and treatment, as well as storage conditions) [5]. To predict the characteristics of CLF droplet ignition and combustion reliably, it is necessary to use the real properties of solid particles and liquids included in the fuel mixture (primarily thermophysical and kinetic characteristics of phase transitions and chemical transformations). An undoubted advantage of modeling is that the component composition fuel suspension, droplet sizes, heat transfer mechanisms and characteristics of heat supply from the external source can be varied in wide ranges. Thus, it is possible to efficiently assess the potential of using promising fuel compositions in practice.
In the research, the numerical simulation was carried out for single droplets (with the initial diameter dp = 1 mm) of five different CLF compositions based on wet FC with MSW: No. 1—FC 100%; No. 2—FC 90% + wood 10%; No. 3—FC 90% + rubber 10%; No. 4—FC 90% + plastic 10%; No. 5—FC 90% + cardboard 10%.
Adding 10% of MSW corresponds to the experimental findings [11,23,28,37] that established that 5–20% of typical combustible MSW added to the CLF composition does not affect the its mechanisms or energy characteristics. At the same time, based on the estimates [38], adding 10–20% typical MSW to the CLF composition will reduce the area of land excluded from agricultural use due to landfill site arrangements by 20–30%. It will also save depleting fossil fuels burned for heat and power generation.

2.2. Experimental Data

The experimental research [23] into the ignition and combustion of single CLF droplets was conducted on two different test installations. The first condition [23] (Figure 1a) created the radiation-supported heating conditions of a suspended droplet in a tubular muffle furnace (Vg ≈ 0) with a temperature varying in the range of 723–1273 K. Using the second condition (Figure 1b), the same fuel droplet was heated convectively with hot air at a velocity of Vg ≈ 3 m/s, the air temperature varied within a range of 723–973 K [23]. The second experimental condition (Figure 1b), in laboratory conditions, simulated the interaction processes of fuel droplets with a stream of hot air, which are the basis of industrial technologies for the flaring of liquid and solid fuels [39].
Carrying out experiments on the demonstrated conditions [23] made it possible to comprehensively analyze the effect of various heating conditions on the ignition characteristics of CLF droplets (Figure 2). The characteristics of ignition under near-critical conditions substantially depend on the patterns of phase transitions and the mechanism of heat and mass transfer for various methods of heat supply to the CLF droplet [40]. Fuel droplets exposed to radiant (Figure 2a) and convective (Figure 2b) heating feature identical patterns of physical and chemical processes. Conditionally the following sequence of basic reaction phases of a CLF droplet and warmed-over air (moving and motionless) can be distinguished: passive warming, near-surface water vaporization, thermolysis of coal and municipal waste (solid combustible components), blending of ignitable gases with the oxidizing medium, gas blend firing and burnout, heating of the solid remainder and finally its heterogeneous ignition and combustion. The main difference between the processes presented in the video frames (in Figure 2a,b) is in the significantly longer passive warming of the fuel droplet and formation of a combustible gas mixture mainly due to the diffusion of volatiles in the oxidizer medium in the vicinity under radiant heating conditions (Figure 2a).
In the experimental research (Figure 2) [23], the ignition-delay times (td) were determined using the commercial Tema Automotive software (Image Systems AB, Sweden) and the built-in Threshold algorithm by the variation of the luminance around the droplet over time [11]. The systematic error of the measuring system in determining the ignition-delay time is less than 3%. Random errors of a series of five repeated measurements under the same experimental conditions are less than 10%. The main experimental values [23] of the average td of droplets (with the initial diameter dp = 1 mm) of CLF with the addition of various MSW at different air temperatures under conditions of dominant radiant and convective heating are presented in Table 3 and Table 4. The calculation results for the developed model will be validated by comparison with these experimental data.

2.3. Mathematical Model

The structure of CLF is a rather complex system. When CLFs are heated, a set of interdependent processes, phase transformations and chemical reactions occurs [41]: passive warming, near-surface water vaporization, thermolysis of coal and municipal waste (solid combustible components), blending of ignitable gases with the oxidizing medium, gas blend firing and burnout, heating of the solid remainder followed by its heterogeneous ignition and combustion. These processes were described in the development of the mathematical model in Ansys Fluent. The discrete particle model (DPM) for composite fuels has been developed (Eulerian–Lagrangian model).
The gas is defined as a stationary medium. The gas phase is described by a stationary Reynolds-averaged Navier–Stokes (RANS) model (1)–(3) [42].
( ρ u ) = 0
( ρ u u ) = p + ( τ ¯ ) + ρ g + F
[ u ( ρ H + p ) ] = ( i h i J i )
where ∇—nabla operator; F —vector of volumetric fluid–particle interaction force, N; H—enthalpy, W; τ ¯ —gas viscous stress, Pa; J—mass flux, kg/(m2·s); h—sensible heat, W; and subscripts i—species.
One of the most important parameters characterizing turbulent motion is turbulent kinetic energy k. The transfer equation for turbulent kinetic energy is derived from the Navier–Stokes and Reynolds equations and is defined as (4):
div ( ρ k u ) = div [ ( μ + μ t σ k ) grad k ] + G k ρ ε
For finding viscous dissipation ε, the transfer equation (k–ε model) is used (5):
div ( ρ ε u ) = div [ ( μ + μ t σ ε ) grad ε ] + C ε 1 ε k G k C ε 2 ρ ε 2 k
where Gk—the generation of turbulence kinetic energy related to the mean velocity gradient and the turbulent model constants Cε1 = 1.44; Cε2 = 1.92; σk = 1; σε = 1.3.
The gas species equation is defined as (6):
( ρ u Y ) = J + R p
where ∇—nabla operator; Rp—rate of production, units vary.
The combusting particle is used as a physical model in the discrete particle model (DPM) to define a single CLF droplet in a gaseous flow. The combusting particle is described by a system of the following laws: heating/cooling, release of volatiles, heterogeneous surface combustion [42]. Meanwhile, at each moment of time there is no temperature gradient inside the combustible particle. It is considered a body with an almost uniform temperature distribution.
Since the model is developed for slurry fuel, it is necessary to use a wet combustion model, so the combusting particle has been extended to include the fraction of the liquid component capable of boiling and vaporizing. The vaporization rate is defined as:
N i = k c ( C i , s C i , )
where Ni—molar flux of vapor, kg∙mol/(m2∙s); kc—mass transfer coefficient, m/s; Ci,s—vapor concentration at the droplet surface, kg∙mol/m3; and Ci,∞—vapor concentration in the bulk gas, kg∙mol/m3.
Dalton’s law determines the vapor concentration near the surface of a fuel droplet:
C i , s = p sat ( T p ) R T p
where R—the universal gas constant; psat—pressure of saturated vapor; and Tp—droplet temperature.
The concentration of vapor in the bulk gas is known from the solution to the transport equation for species i for non-premixed or partially premixed combustion calculations:
C i , = X i p R T
where Xi—the local bulk mole fraction of species i; p—the local absolute pressure; and T—the local bulk temperature in the gas.
The Rantz and Marshall correlation was used to estimate the mass transfer coefficient kc from the Sherwood number [42]:
Sh = k c d p D i , m = 2.0 + 0.6 Re d 1 2 Sc 1 3
where Sc—the Schmidt number, Sc = μ ρ D i , m .
When moisture evaporates from a droplet, its mass decreases according to the equation:
m p ( t + Δ t ) = m p ( t ) N i A p M w , i Δ t
where Mw,i—molecular weight of species i, kg/(mol·K).
The rate of convective boiling of the liquid contained in the droplet after reaching the boiling point (Tbp):
d ( d p ) d t = 4 k p p c p , d p ( 1 + 0.23 Re d ) ln [ 1 + c p , ( T T p ) h fg ]
The motion equation is described using Newton’s second law of motion:
d u p d t = F D ( u u p ) + g ( ρ p ρ ) ρ p + F a
where Fa—additional acceleration, N/kg; FD (u − up)—drag force, N/kg; and FD—coefficient in drag law, 1/s.
Coefficient in drag law FD for a spherical droplet is defined as:
F D = 18 μ ρ p d p 2 C D Re 24
where CD—drag coefficient for spherical droplet.
Relative Reynolds number Re is defined as:
Re = ρ d p | u p u | μ
The heating of the combusting particle is described by the energy equation:
m p c p d T p d t = h A p ( T T p ) + d m p d t h f g + A p ε p σ ( θ R 4 T p 4 )
m p c p d T p d t = h A p ( T T p ) f h d m p d t H reac + A p ε p σ ( θ R 4 T p 4 )
Energy equations of particles (16) and (17) are solved analytically, by assuming that the temperature and mass of the particle do not change significantly between time steps [42]:
T p ( t + Δ t ) = α p + [ T p ( t ) α p ] e β p Δ t
Coefficients of αp and βp in equation (18) are defined as:
α p = h A p T + d m p d t h f g + A p ε p σ θ R 4 h A p + A p ε p σ T p 3
β p = A p ( h + ε p σ T p 3 ) m p c p
where hfg—latent heat, J/kg; Hreac—heat released by the surface reaction, J/kg; fh—particle fraction which absorbs heat of reaction; εp—particle emissivity; σ—Stefan–Boltzmann constant, W/(m2·K4); and θR—radiation temperature, K.
The Rantz and Marshall correlation was also used to estimate the heat transfer coefficient in accordance with the main provisions [43,44], since the CLF droplet is considered as a system with an almost uniform temperature distribution. Convective heat transfer coefficient h is defined from the Nusselt number (21) which is a function of particle Reynolds number (15) and gas Prandtl number:
Nu = h d p k = 2.0 + 0.6 Re 1 / 2 Pr 1 / 3
Pr = c p μ k
The devolatilization process (23) releases volatiles (CαHβOγNδ) and char (Cs) [42]:
Solid fuel C α H β O γ N δ + C s
The rate of devolatilization is determined as a constant value according to the equation [45,46]:
1 f v , 0 ( 1 f w , 0 ) m p , 0 d m p d t = A 0
where fv,0—initial mass fraction of volatiles; fw,0—initial mass fraction of evaporating material; mp,0—initial particle mass, kg; and A0—rate constant, s−1.
Note that the volatilization process occurs when the particles of coal reach the evaporation temperature (in our case vaporization temperature is equal to 400 K [47]). Particle swelling effects during devolatilization are ignored, which is consistent with experimental observations [28,41,48]. The change of particle temperature is determined from the energy equations of particles (16) and (17) governed by conductive, convective and radiative heat transfer [42].
The multiple surface reactions model is used for the description of the reacting particle. Any number of chemical reactions can be described using the multiple surface reactions model [42]. The scheme for a reacting particle in the multiple surface reactions model is shown in Figure 3.
The rate of chemical reactions:
R ¯ j , r = A p η r Y j R j , r
R j , r = R kin , r ( p n R j , r D 0 , r ) N r
where R ¯ j , r —rate of particle surface species depletion, kg/s; Ap—particle surface area, m2; ηr—effectiveness factor; Rj,r—rate of particle surface species reaction per unit area, kg/(m2·s); pn—bulk partial pressure of the gas phase species, Pa; Rkin,r—kinetic rate of reaction, units vary; D0,r—diffusion rate coefficient for reaction; and Nr—apparent order of reaction.
The diffusion rate coefficient D0,r is defined as:
D 0 , r = C 1 , r [ ( T p + T ) / 2 ] 0.75 d p
The kinetic rate of multiple reactions is defined as:
R kin , r = A r T p β r e ( E r / R T p ) ( p r , d ) N r n = 1 n max p n N r , n
where Ar—pre-exponential factor; βr—temperature exponent; Er—activation energy for reaction, J/(mol·K); pr,d—bulk partial pressure of the diffusion-limited species, Pa; and subscripts: n—number of gas specie.
The main kinetic parameters of chemical reactions used in the modeling are presented in Table 5.
The thermophysical and transport properties of multi-component CLF droplets, which are a heterogeneous mixture of coal, water and MSW, were set as effective values, which were determined analytically using the method [70] based on the data presented in Table 3 and Table 4. The obtained values are presented in Table 6.
For the developed model of CLT droplet ignition under different heating conditions, a structured rectangular grid 30 × 30 mm in size was generated using ANSYS Fluent [42]. To assess the influence of the grid element sizes on the simulation results, a series of calculations was carried out with a gradual increase in the number of elements. The calculations were carried out for grids consisting of 5625, 14,400 and 360,000 elements (Table 7). The grid convergence index (GCI) was calculated according to the recommended procedure for the estimation of discretization errors [71,72]. The number of elements and the characteristic size of the computational grid were considered optimal when the corresponding convergence criteria were met (GCI ≤ 0.05%). Grid 2 satisfies this condition (Table 7).
For the computational model, the grid was built in such a way that the value of the maximum size of the grid element was 0.25 mm (14,400 elements in total, 14,641 nodes). In this area of the problem’s solution (Figure 4), the ongoing physical and chemical processes were controlled.
A single droplet injection system was used to generate a CLF droplet with the initial temperature Tp,0 = 300 K and size dp = 1 mm. It was placed in the geometric median of the solution domain in Figure 4 (X-position = 0.015 m; Y-position = 0.015 m). At the initial point of time, a uniform distribution of the oxidizer temperature Tg was set. The boundaries of the solution domain (furnace chamber walls) were described by the first type (Dirichlet) of boundary conditions. The temperatures at the solution domain boundaries were set as constant and equal to Tg according to the experimental data. In the modeling of the composite fuel droplet ignition under the convective heating conditions, the velocity of heated air Vg was 3 m/s. The oxidizer (O2) concentration was assumed to be equal to the mass content of oxidizer in the air φO2 = 0.21.
Verification of the developed mathematical model and assessment of the reliability of the results of numerical simulation were carried out by a standard method for checking the conservatism of the difference scheme used. The step along the temporal and spatial coordinates was chosen by the method of half division until the constancy of the controlled characteristics was achieved. The error of fulfillment of the law of conservation of energy in the area of solution of the ignition problem was calculated at each time step. In the system under consideration, when calculating the ignition characteristics with a step in the spatial coordinate Δr = 250 μm and in time Δt = 25 μs a relatively small error in the performance of the energy balance is achieved (the integral value did not exceed 2.5%).
The target characteristic of the simulated process was the ignition-delay time, which was determined as follows. For td we took [73] the period of time from the moment the heating of the fuel droplet began (the beginning of the calculation) until the moment it reached the ignition temperature Td = 723 K. Figure 5 shows a typical graphical illustration of the determination of the ignition-delay time of the fuel droplet with an initial diameter dp = 1 mm.

3. Results and Discussion

In the numerical simulation, some constants from the source data were not known precisely, thus their values were assumed, on the basis of empirical data, to match the numerical simulation results with the experiment. To establish the main process characteristic (ignition-delay time) as a function of a group of source data, the rates of volatile release from the solid part of the fuel mixture, reaction heat fractions absorbed by the solid and the volatile component fraction/combustible fraction ratios were varied in allowable ranges in Ansys Fluent. Moreover, temperature trends at different heating source temperatures were obtained. The results are presented in Figure 6, Figure 7, Figure 8 and Figure 9.
As seen in Figure 6, changing the devolatilization constant of the combusting particle has a significant impact on the ignition-delay time only when it is varied between 0 and 5 s−1. The ignition-delay time in this condition is an interval from the calculation initiation until the moment the droplet reaches the ignition temperature. Thus, it includes the gas-phase ignition time and the time the solid residue is heated until the ignition temperature (heterogeneous ignition):
t d = t g f ( D C ) + t h g ( T g )
where tgf—gas-phase ignition-delay time of the CLF (depending on the devolatilization rate), s; and thg—heterogeneous ignition-delay time of the CLF (depending on the ambient air temperature), s.
Thus, the rate of devolatilization affects primarily the fuel gas-phase ignition time (at a constant devolatilization rate, the gas-phase ignition time is constant). Due to the low content of volatiles in the fuel, the time of the gas-phase ignition of the fuel is much shorter compared to the time of heterogeneous ignition. Despite this fact, the influence of this parameter becomes more considerable with an increase in the surrounding oxidizer temperature. The reason for that is that a higher temperature reduces the heterogeneous ignition-delay time, thus increasing the significance of the gas-phase ignition delay which is constant at a constant devolatilization rate.
As the CLF droplet interacts with the heated air, exothermic and endothermic chemical reactions evolve on the surface. Some of the heat they release is absorbed by the fuel solid residue. The higher this value, the faster the solid residue is heated and the lower the ignition-delay time. With the growing ambient temperature, the effect of the absorbed energy from the chemical interaction on the ignition considerably weakens, since the bulk of the energy will come from the radiant and convective heat exchange with the ambient air (Figure 7).
The curves of the CLF ignition-delay times vs. the volatile component fraction/combustible fraction ratio of the combusting particle (Figure 8) are close to linear. The total ignition-delay time diminishes as the combustible fraction decreases. This is because the lower the combustible fraction in the fuel composition, the lower the heterogeneous ignition time which makes up most of the total ignition time of CLF. Moreover, the effect of this ratio is significantly reduced by an increase in the surrounding oxidizer temperature. This pattern is generally confirmed by the experimental research [23]: the higher the temperature, the less influence the heat exchange in the vicinity of the fuel droplet has on the ignition-delay time.
Thus, for better reliability in the prediction of the CLF ignition characteristics with numerical simulation, the data for the volatile component fraction/combustible fraction ratio of the combusting particle need to be determined precisely as source values, whereas the devolatilization constant and reaction heat fractions absorbed by the solid of the combusting particle do not require high accuracy and can be taken from reference materials.
Using the developed mathematical model, the temperature trends of the CLF droplet were obtained during heating, ignition and combustion (Figure 9a) at different temperatures of the heating source (723 K, 973 K and 1273 K). It has been established that the moment at which the maximum temperature is reached shifts to the left with an increase in the temperature of the heating source, which is mainly due to the faster occurrence of a set of physicochemical reactions. The dynamics of the change in the temperature of CLF droplets is of great practical importance, in particular, for the analysis of the energy characteristics of the fuel combustion process. [10,11]. In Figure 9b theoretical and experimental temperature trends are compared. It should be noted that there is a satisfactory qualitative agreement between theory and experiment. At the same time, the quantitative differences are significant and reach more than 20–25%, which is mainly due to the assumptions made in the basic mathematical model. Firstly, constant values of evaporation and devolatilization rates were used in the numerical study. Obviously, under real conditions, the rates of evaporation and devolatilization are not constant, but depend on temperature and change during heating. Second, the model did not take into account the non-uniformity of the temperature distribution inside the CLF droplet and its vicinity, which is the root cause of the change in the rates of physicochemical transformations. Thirdly, the developed model makes it possible to obtain the temperature values directly inside the droplet. The experimental data curve illustrates the temperature in the vicinity of the CLF drop, where the gas mixture burns out, which is formed during the evaporation of the liquid combustible component of the fuel (if present) and during the thermal decomposition of solid components (FC and MSW) [41].
Figure 10 presents the curves of the ignition-delay times vs. oxidizer temperature for CLF with different component compositions. The curves of the ignition-delay times vs. oxidizer temperature are nonlinear. This is conditioned by the fact that, with an increase in temperature, the rates of the thermolysis of organic fuel components, water vaporization, combustion of ignitable gases, and heterogeneous firing of solid remainder change exponentially. This dependence of the ignition-delay times on the ambient temperature was also recorded in the experimental research [23].
Droplets of fuel composition No. 1 (FC 100%) are characterized by maximum ignition-delay times which are due to the high humidity of the initial fuel composition. Induction period of composition No. 1 (Figure 10a) has a longer duration compared with other fuels due to the removal of more energy for the water evaporation. Divergence between the experimental data and the results of numerical simulation for composition No. 1 is 2–43% under radiant heating conditions (Vg ≈ 0) and 5–32% under convective heating conditions (Vg ≈ 3 m/s). Differences in ignition-delay times for composition No. 2 (Figure 10b) (FC 90% + wood 10%) over the entire range of temperature variation are 2–48% and 5–29% (with radiant and convective heating, respectively); for other compositions, similar characteristics are as follows: No. 3 (Figure 10c) (FC 90% + rubber 10%)—9–44% and 7–32%; No. 4 (Figure 10d) (FC 90% + plastic 10%)—1–48% and 1–27%; No. 5 (Figure 10e) (FC 90% + cardboard 10%)—10–54% and 3–23%. The last two compositions with the addition of plastic and cardboard (No. 4 and No. 5) are characterized by the most intense ignition of the droplet, which correlates with the experimental data. This is due to the higher content of volatile matter in the fuel composition and its relatively low moisture content. The performed analysis allows us to conclude that the numerical simulation results are in satisfactory agreement with the experimental data.
It should be noted that the numerical simulation of the processes under study was carried out for somewhat idealized fuel compositions without impurities, which, obviously, is not carried out in practice. We believe that the main differences in the results are due both to the heat losses not taken into account in the model, which are typical during the experiment, and to the occurrence of additional chemical reactions with impurities contained in the real fuel. Impurities can be caused by the content of flotation reagents in coal processing waste (oil refinery products and petrochemical waste). In addition, the combustible MSWs (rubber, wood, plastic) used are chemically complex organic substances, during the combustion of which intermediate components can interact with other fuel components.
The developed mathematical model is most sensitive to a change in the volatile/combustible fraction ratio of the combusting particle. Therefore, for better congruence between the numerical simulation results and experimental data, it is a relevant task to build up an experimental database with reliable experimental values of such constants for different CLF compositions.
An important result of the conducted research is the development of a model that has no limitations on the component composition of fuel slurries and heating conditions. Thus, the model can be applied when the existing power generating facilities are switched to composite fuel. This enables one to evaluate the ignition and combustion characteristics of the fuel, whose properties may vary greatly at different industrial enterprises or within one enterprise when MSW with a different content of combustible components is used. The model can be used to predict the main integral ignition characteristics of CLF droplets in the operation of current and prospective steam and hot-water boilers. These include, in particular, the ignition-delay times, CLF droplet temperature, fuel component composition changes during heating, unsteady release of volatiles, etc. In the future, the reliability of the numerical simulation results will be improved by integrating user-defined functions (UDF) in the algorithm developed in Ansys Fluent to describe a group of constants, e.g., thermophysical characteristics, as functions of temperature or other external and internal factors.
It should be noted that an increase in the reliability of the results of numerical simulation (Figure 10) can be achieved by taking into account the effect of temperature gradient inside the droplet and in its vicinity. The inhomogeneity of the temperature fields leads to an increase in the duration of the fuel droplet heating period until the moment of ignition in comparison with the previously described idealized ignition problem. However, the description of transient heat transfer processes requires large computing power of the PC.

4. Effect of Temperature Gradient Inside the Droplet and in Its Vicinity

This section presents the first results on the modernization of the developed mathematical model for a more reliable description of the experimental data. Compared to the base model (1)–(28), the advanced version of the model additionally takes into account the non-uniform temperature distribution inside the CLF droplet according to the heat transfer equation:
T p t = a p ( 2 T p r 2 + 2 r T p r )
where r—droplet radius in spherical coordinates, m; Tp—droplet temperature, K; and ap—thermal diffusivity of CLF, m2/s.
The complexity of the mathematical model made it possible to construct typical temperature fields for solving the problem of ignition of a CLF droplet for qualitative comparison with videograms of experimental studies [23] (Figure 11). The first case (Figure 11a) was radiant heating (Vg ≈ 0). The second case (Figure 11b) was convective heating (Vg ≈ 3 m/s). As can be seen, there is a qualitative agreement between experimental data and simulation results in the stages of industrial waste-derived fuel droplet combustion: passive warming, near-surface water vaporization, thermolysis of coal and municipal waste (solid combustible components), blending of ignitable gases with the oxidizing medium, gas blend firing and burnout, and heating of the solid remainder followed by its heterogeneous ignition and combustion.
Using the advanced mathematical model, ignition-delay times td are determined depending on the temperature of the heating source Tg. Satisfactory agreement between experiment and theory was obtained (Figure 12). In this case, the difference in the results for td over the entire range of temperatures under study does not exceed 19% both under conditions of radiant heating (Figure 12a) and convective heating (Figure 12b), which does not exceed the confidence limit for the obtained experimental data [23].
It should be noted that when the non-uniform temperature distribution inside a single CLF droplet and in its vicinity is taken into account, the PC calculation times increase by a factor of 5–10. On the scale of the steam boiler, there will be an exponential increase in the PC calculation times. On the scale of the steam boiler, there will be an exponential increase in the PC calculation times. The relatively simple modernization performed on the mathematical model, on the one hand, leads to a satisfactory agreement between the results of numerical simulation and experimental data, but, on the other hand, requires large computing power. The most rational approach in practical applications is the development of a mathematical model with maximum simplifications but with acceptable reproducibility of experimental data.

5. Conclusions

Based on results of previous experimental studies, a mathematical model was developed for the ignition and combustion of a single composite liquid fuel droplet (a mixture of coal processing waste in a wet state with combustible municipal solid waste). The model considers the interdependent sequence of the main chemical transformations and physical interactions of the suspended CLF droplet with the surrounding air environment (passive warming, near-surface water vaporization, thermolysis of coal and municipal waste (solid combustible components), blending of ignitable gases with the oxidizing medium, gas blend firing and burnout, and heating of the solid remainder followed by its heterogeneous ignition and combustion with an ash residue formation) as a result of combined radiant and convective heat transfer, the heating source temperature for which was varied within the ranges of 723–1273 K and 723–973 K, respectively. The numerical simulation of combustion characteristics was carried out in the Ansys Fluent commercial software for five different CLF compositions. Comparison of the ignition-delay times for single droplets obtained by numerical simulation with experimental data showed their good agreement for mathematical model configuration, considering the temperature gradient inside the fuel droplet and in its vicinity. The results of numerical simulation of a CLF droplet in a heated air flow differ from the experimental data by no more than 19%, both under conditions of radiant heating and convective heating, which does not exceed the confidence limit for the obtained experimental data. The developed mathematical model of the ignition and combustion process of a single CLF droplet makes it possible to predict the main characteristics with good reliability. The mathematical model is most sensitive to changes in the ratio of volatile/combustible fractions of the combusting particle, therefore, an urgent task in the improvement of the correspondence of numerical simulation results with experimental data is to compile an information base with reliable experimental values of such constants for different CLF compositions. It can also be noted that the predictive mechanism developed in Ansys Fluent can significantly reduce the time and material costs of subsequent research.

Author Contributions

Conceptualization, D.G.; data curation, A.R.; formal analysis, D.A.; funding acquisition, D.G.; investigation, K.P.; methodology, D.A.; project administration, K.P.; writing—original draft, K.P.; writing—review and editing, D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Foundation for Basic Research, National Council of Brazil for Scientific and Technological Development and Ministry of Science & Technology (Government of India) grant number [19-53-80019].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

CFcombustible fraction
CLFcomposite liquid fuel
DCdevolatilization constant
DPMdiscrete particle model
FCfilter cake
MSWmunicipal solid waste
RANSReynolds-averaged Navier-Stokes equations
TPPthermal power plant
VFvolatile fraction

Nomenclature

Adash content, %
apthermal diffusivity of CLF, m2/s
Apsurface area of the droplet, m2
Arpre-exponential factor
A0rate constant, s−1
CDdrag coefficient for spherical droplet
Cdaf, Hdaf, Ndaf, Odaf, Sdaffraction of carbon, hydrogen, nitrogen, oxygen, sulfur in the sample converted to a dry ash free state, %
Cd,b, Cd,sconcentration of the diffusion-limited species
Ci,svapor concentration at the droplet surface, kg·mol/m3
Ci,∞vapor concentration in the bulk gas, kg·mol/m3
Ckconcentrations of all other species
cpparticle (droplet) heat capacity, J/(kg·K)
cp,∞gas heat capacity, J/(kg∙K)
C1,rmolar concentration of species j in reactions, mol
Di,mdiffusion coefficient of vapor in the bulk, m2/s
dpparticle (droplet) diameter, m
D0,rdiffusion rate coefficient for reaction
Eractivation energy for reaction, J/(mol·K)
F vector of volumetric fluid / particle interaction force, N
Faadditional acceleration, N/kg
FDcoefficient in drag law, 1/s
fhparticle fraction which absorbs heat of reaction
fv,0initial mass fraction of volatiles
fw,0initial mass fraction of evaporating material
g gravity acceleration, m/s2
Gkgeneration of turbulence kinetic energy related to the mean velocity gradient and the turbulent model constants
Henthalpy, W
hconvective heat transfer coefficient, W/(m2·K)
hfglatent heat, J/kg
hisensible heat, W
Hreacheat released by the surface reaction, J/kg
J mass flux, kg/(m2·s)
kturbulent kinetic energy, m2/s2
kcmass transfer coefficient, m/s
kthermal conductivity of the gas, W/(m∙K)
mpparticle (droplet) mass, kg
mp,0initial particle (droplet) mass, kg
Mw,imolecular weight of species i, kg/mol·K
Nimolar flux of vapor, kg·mol/(m2∙s)
Nrapparent order of reaction
NuNusselt number
pgas local absolute pressure, Pa
pnbulk partial pressure of the gas phase species, Pa
PrPrandtl number
pr,dbulk partial pressure of the diffusion-limited species, Pa
psatsaturated vapor pressure, Pa
Qheat of combustion, J/kg
Qas,Vhigher heating value, J/kg
rdroplet radius in spherical coordinates, m
Rthe universal gas constant, J/(K·mol)
Rprate of production, units vary
Rerelative Reynolds number
R ¯ j , r rate of particle surface species depletion, kg/s
Rj,rrate of particle surface species reaction per unit area, kg/(m2·s)
Rkin,rkinetic rate of reaction, units vary
ScSchmidt number
ShSherwood number
ttime, s
Tbpboiling point, K
TdCLF ignition temperature, K
tdignition delay time, s
Tgoxidizer temperature, K
tgfgas-phase ignition-delay time, s
thgheterogeneous ignition-delay time of CLF, s
Tpdroplet temperature, K
Tp,0initial particle (droplet) temperature, K
Tvapvaporization temperature, K
Tgas temperature, K
u velocity vector, m/s
ugas velocity, m/s
upparticle velocity, m/s
Vdafvolatile content, %
Vgoxidizer velocity, m/s
Wamoisture content, %
Xilocal bulk mole fraction of species i
Ymass fraction of solid fuel
Yjmass fraction of surface species j in the particle
Greek symbols
βrtemperature exponent
Δrstep on spatial coordinate, m
Δttime step, s
εturbulent dissipation rate, m2/s3
εpparticle emissivity
ηreffectiveness factor
θRradiation temperature, K
µgas dynamic viscosity, Pa·s
µtturbulent dynamic viscosity, Pa·s
ρgas density, kg/m3
ρpparticle (droplet) density, kg/m3
σStefan-Boltzmann constant, W/(m2·K4)
σkturbulent Prandtl number for turbulent kinetic energy
σεturbulent Prandtl number for turbulent dissipation rate
τ ¯ gas viscous stress, Pa
φO2mass content of oxygen in the air
Subscripts
ispecies
nnumber of gas phase specie

References

  1. Khan, H.H.; Malik, M.N.; Zafar, R.; Goni, F.A.; Chofreh, A.G.; Klemeš, J.J.; Alotaibi, Y. Challenges for Sustainable Smart City Development: A Conceptual Framework. Sustain. Dev. 2020, 28, 1507–1518. [Google Scholar] [CrossRef]
  2. IEA World Energy Outlook 2020. Available online: https://iea.blob.core.windows.net/assets/a72d8abf-de08-4385-8711-b8a062d6124a/WEO2020.pdf (accessed on 1 March 2021).
  3. Global Information Inc. Global Industrial Waste Management Market-2019–2026; Global Information Inc.: Tokyo, Japan, 2020. [Google Scholar]
  4. Gupta, A.; Paranjape, N. Industrial Solid Waste Management Market 2020–2026; Global Market Insights Inc.: Selbyville, DE, USA, 2020. [Google Scholar]
  5. Kaza, S.; Yao, L.C.; Bhada-Tata, P.; Van Woerden, F. What a Waste 2.0: A Global Snapshot of Solid Waste Management to 2050; World Bank: Washington, DC, USA, 2018; ISBN 978-1-4648-1329-0. [Google Scholar]
  6. Sabour, M.R.; Alam, E.; Hatami, A.M. Global Trends and Status in Landfilling Research: A Systematic Analysis. J. Mater. Cycles Waste Manag. 2020, 22, 711–723. [Google Scholar] [CrossRef]
  7. Chen, D.M.C.; Bodirsky, B.L.; Krueger, T.; Mishra, A.; Popp, A. The World’s Growing Municipal Solid Waste: Trends and Impacts. Environ. Res. Lett. 2020, 15, 074021. [Google Scholar] [CrossRef]
  8. Xin, C.; Zhang, T.; Tsai, S.-B.; Zhai, Y.-M.; Wang, J. An Empirical Study on Greenhouse Gas Emission Calculations under Different Municipal Solid Waste Management Strategies. Appl. Sci. 2020, 10, 1673. [Google Scholar] [CrossRef] [Green Version]
  9. Han, Z.; Wang, S.; Zhao, J.; Hu, X.; Fei, Y.; Xu, M. Identification of Nitrogen-Sources in an Aquifer beneath a Municipal Solid Waste Landfill in the Vicinity of Multiple Pollutant Sources. J. Environ. Manag. 2020, 268, 110661. [Google Scholar] [CrossRef] [PubMed]
  10. Foster, W.; Azimov, U.; Gauthier-Maradei, P.; Molano, L.C.; Combrinck, M.; Munoz, J.; Esteves, J.J.; Patino, L. Waste-to-Energy Conversion Technologies in the UK: Processes and Barriers–A Review. Renew. Sustain. Energy Rev. 2021, 135, 110226. [Google Scholar] [CrossRef]
  11. Glushkov, D.O.; Strizhak, P.A. Ignition of Composite Liquid Fuel Droplets Based on Coal and Oil Processing Waste by Heated Air Flow. J. Clean. Prod. 2017, 165, 1445–1461. [Google Scholar] [CrossRef]
  12. Staroń, A.; Kowalski, Z.; Staroń, P.; Banach, M. Studies on CWL with Glycerol for Combustion Process. Environ. Sci. Pollut. Res. 2019, 26, 2835–2844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhao, Z.; Wang, R.; Ge, L.; Wu, J.; Yin, Q.; Wang, C. Energy Utilization of Coal-Coking Wastes via Coal Slurry Preparation: The Characteristics of Slurrying, Combustion, and Pollutant Emission. Energy 2019, 168, 609–618. [Google Scholar] [CrossRef]
  14. Aliukov, S.; Osintsev, K. Mathematical Modeling of Coal Dust Screening by Means of Sieve Analysis and Coal Dust Combustion Based on New Methods of Piece-linear Function Approximation. Appl. Sci. 2021, 11, 1609. [Google Scholar] [CrossRef]
  15. Sun, W.; Zhong, W.; Zhang, J.; Echekki, T. Large Eddy Simulation on the Effects of Coal Particles Size on Turbulent Combustion Characteristics and NOx Formation inside a Corner-Fired Furnace. J. Energy Resour. Technol. 2021, 143, 082302. [Google Scholar] [CrossRef]
  16. Gu, T.; Yin, C.; Ma, W.; Chen, G. Municipal Solid Waste Incineration in a Packed Bed: A Comprehensive Modeling Study with Experimental Validation. Appl. Energy 2019, 247, 127–139. [Google Scholar] [CrossRef]
  17. Janajreh, I.; Adeyemi, I.; Elagroudy, S. Gasification Feasibility of Polyethylene, Polypropylene, Polystyrene Waste and Their Mixture: Experimental Studies and Modeling. Sustain. Energy Technol. Assess. 2020, 39, 100684. [Google Scholar] [CrossRef]
  18. Bahari, A.; Atashkari, K.; Mahmoudimehr, J. Multi-Objective Optimization of a Municipal Solid Waste Gasifier. Biomass Convers. Biorefinery 2020, 11, 1703–1718. [Google Scholar] [CrossRef]
  19. Somwangthanaroj, S.; Fukuda, S. Investigation of Multi-Fuel Combustion Behavior and Synergy Effect Using Improved Steady-State Discrete Particle Model Simulation. Int. J. Energy Res. 2020, 45, 17291–17301. [Google Scholar] [CrossRef]
  20. Zhuo, Y.; Shen, Y. Three-Dimensional Transient Modelling of Coal and Coke Co-Combustion in the Dynamic Raceway of Ironmaking Blast Furnaces. Appl. Energy 2020, 261, 114456. [Google Scholar] [CrossRef]
  21. Tokmurzin, D.; Adair, D. Development of Euler-Lagrangian Simulation of a Circulating Fluidized Bed Reactor for Coal Gasification. Eurasian Chem. J. 2019, 21, 45–49. [Google Scholar] [CrossRef]
  22. Yang, R.; Ma, C.; Chen, G.; Cheng, Z.; Yan, B.; Mansour, M. Study on NOx Emission during Corn Straw/Sewage Sludge Co-Combustion: Experiments and Modelling. Fuel 2021, 285, 119208. [Google Scholar] [CrossRef]
  23. Glushkov, D.O.; Paushkina, K.K.; Shabardin, D.P.; Strizhak, P.A. Environmental Aspects of Converting Municipal Solid Waste into Energy as Part of Composite Fuels. J. Clean. Prod. 2018, 201, 1029–1042. [Google Scholar] [CrossRef]
  24. Glushkov, D.O.; Feoktistov, D.V.; Kuznetsov, G.V.; Batishcheva, K.A.; Kudelova, T.; Paushkina, K.K. Conditions and Characteristics of Droplets Breakup for Industrial Waste-Derived Fuel Suspensions Ignited in High-Temperature Air. Fuel 2020, 265, 116915. [Google Scholar] [CrossRef]
  25. Nyashina, G.S.; Legros, J.C.; Strizhak, P.A. Environmental Potential of Using Coal-Processing Waste as the Primary and Secondary Fuel for Energy Providers. Energies 2017, 10, 405. [Google Scholar] [CrossRef] [Green Version]
  26. Nyashina, G.S.; Kuznetsov, G.V.; Strizhak, P.A. Energy Efficiency and Environmental Aspects of the Combustion of Coal-Water Slurries with and without Petrochemicals. J. Clean. Prod. 2018, 172, 1730–1738. [Google Scholar] [CrossRef]
  27. Nyashina, G.S.; Kurgankina, M.A.; Strizhak, P.A. Environmental, Economic and Energetic Benefits of Using Coal and Oil Processing Waste Instead of Coal to Produce the Same Amount of Energy. Energy Convers. Manag. 2018, 174, 175–187. [Google Scholar] [CrossRef]
  28. Glushkov, D.O.; Paushkina, K.K.; Shabardin, D.P.; Strizhak, P.A.; Gutareva, N.Y. Municipal Solid Waste Recycling by Burning It as Part of Composite Fuel with Energy Generation. J. Environ. Manag. 2019, 231, 896–904. [Google Scholar] [CrossRef] [PubMed]
  29. Moriguchi, Y.; Hashimoto, S. Material Flow Analysis and Waste Management. In Taking Stock of Industrial Ecology; Springer International Publishing: Cham, Switzerland, 2016; pp. 247–262. ISBN 9783319205717. [Google Scholar]
  30. Edjabou, M.E.; Martín-Fernández, J.A.; Scheutz, C.; Astrup, T.F. Statistical Analysis of Solid Waste Composition Data: Arithmetic Mean, Standard Deviation and Correlation Coefficients. Waste Manag. 2017, 69, 13–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Faitli, J.; Magyar, T.; Erdélyi, A.; Murányi, A. Characterization of Thermal Properties of Municipal Solid Waste Landfills. Waste Manag. 2015, 36, 213–221. [Google Scholar] [CrossRef] [PubMed]
  32. Birgen, C.; Magnanelli, E.; Carlsson, P.; Skreiberg, Ø.; Mosby, J.; Becidan, M. Machine Learning Based Modelling for Lower Heating Value Prediction of Municipal Solid Waste. Fuel 2021, 283, 118906. [Google Scholar] [CrossRef]
  33. Wang, D.; Tang, Y.T.; He, J.; Yang, F.; Robinson, D. Generalized Models to Predict the Lower Heating Value (LHV) of Municipal Solid Waste (MSW). Energy 2021, 216, 119279. [Google Scholar] [CrossRef]
  34. Tabakaev, R.; Shanenkov, I.; Kazakov, A.; Zavorin, A. Thermal Processing of Biomass into High-Calorific Solid Composite Fuel. J. Anal. Appl. Pyrolysis 2017, 124, 94–102. [Google Scholar] [CrossRef]
  35. Ilinykh, G.V. Evaluation of Thermotechnical Properties of Solid Waste from Their Morphological Composition. Vestn. Permsk. Natsionalnogo Issledovatelskogo Politekh. Univ. Urban. 2013, 11, 125–137. [Google Scholar]
  36. Glushkov, D.O.; Strizhak, P.A.; Chernetskii, M.Y. Organic Coal-Water Fuel: Problems and Advances (Review). Therm. Eng. 2016, 63, 707–717. [Google Scholar] [CrossRef]
  37. Glushkov, D.O.; Shabardin, D.P.; Strizhak, P.A.; Vershinina, K.Y. Influence of Organic Coal-Water Fuel Composition on the Characteristics of Sustainable Droplet Ignition. Fuel Process. Technol. 2016, 143, 60–68. [Google Scholar] [CrossRef]
  38. Dmitrienko, M.A.; Strizhak, P.A.; Tsygankova, Y.S. Technoeconomic Analysis of Prospects of Use of Organic Coal-Water Fuels of Various Component Compositions. Chem. Pet. Eng. 2017, 53, 195–202. [Google Scholar] [CrossRef]
  39. Stöllinger, M.; Naud, B.; Roekaerts, D.; Beishuizen, N.; Heinz, S. PDF Modeling and Simulations of Pulverized Coal Combustion-Part 1: Theory and Modeling. Combust. Flame 2013, 160, 384–395. [Google Scholar] [CrossRef]
  40. Lyulin, Y.V.; Feoktistov, D.V.; Afanas’ev, I.A.; Chachilo, E.S.; Kabov, O.A.; Kuznetsov, G.V. Measuring the Rate of Local Evaporation from the Liquid Surface under the Action of Gas Flow. Tech. Phys. Lett. 2015, 41, 665–667. [Google Scholar] [CrossRef]
  41. Glushkov, D.O.; Paushkina, K.K.; Shabardin, D.P. Co-Combustion of Coal Processing Waste, Oil Refining Waste and Municipal Solid Waste: Mechanism, Characteristics, Emissions. Chemosphere 2020, 240, 124892. [Google Scholar] [CrossRef]
  42. ANSYS Academic Research ANSYS Fluent Theory Guide. In ANSYS Help System; ANSYS, Inc.: Canonsburg, PA, USA, 2018.
  43. Ranz, W.E.; Marshall, W.R. Evaporation From Drops–Part I. Chem. Eng. Prog. 1952, 48, 141–146. [Google Scholar]
  44. Aissa, A.; Abdelouahab, M.; Noureddine, A.; Elganaoui, M.; Pateyron, B. Ranz and Marshall Correlations Limits on Heat Flow between a Sphere and Its Surrounding Gas at High Temperature. Therm. Sci. 2015, 19, 1521–1528. [Google Scholar] [CrossRef]
  45. Baum, M.M.; Street, P.J. Predicting the Combustion Behaviour of Coal Particles. Combust. Sci. Technol. 1971, 3, 231–243. [Google Scholar] [CrossRef]
  46. Wijayanta, A.T.; Alam, M.S.; Nakaso, K.; Fukai, J.; Kunitomo, K.; Shimizu, M. Combustibility of Biochar Injected into the Raceway of a Blast Furnace. Fuel Process. Technol. 2014, 117, 53–59. [Google Scholar] [CrossRef]
  47. Ejeh, C.; Afgan, I.; AlMansob, H.; Brantson, E.; Fekala, J.; Odiator, M.; Stanley, P.; Anumah, P.; Onyekperem, C.; Boah, E. Computational Fluid Dynamics for Ameliorating Oil Recovery Using Silicon-Based Nanofluids and Ethanol in Oil-Wet Reservoirs. Energy Rep. 2020, 6, 3023–3035. [Google Scholar] [CrossRef]
  48. Glushkov, D.O.; Kuznetsov, G.V.; Paushkina, K.K.; Shabardin, D.P. The Main Elements of a Strategy for Combined Utilization of Industrial and Municipal Waste from Neighboring Regions by Burning It as Part of Composite Fuels. Energies 2018, 11, 2534. [Google Scholar] [CrossRef]
  49. Taghiyeva, S.F.; Jafarov, R.P.; Kasimov, A.A.; Qadji-Kasumov, V.S.; Rustamov, M.I. Kinetic Model of Hydrogenation of Carbon Dioxide on the Catalyst Fe, Ni/γ-Al2O3. PPOR 2017, 18, 62–68. [Google Scholar]
  50. Wickramaarachchi, W.A.M.K.P.; Narayana, M. Pyrolysis of Single Biomass Particle Using Three-Dimensional Computational Fluid Dynamics Modelling. Renew. Energy 2020, 146, 1153–1165. [Google Scholar] [CrossRef]
  51. Wei, X.; Zhong, H.; Yang, Q.; Yao, E.; Zhang, Y.; Zou, H. Studying the Mechanisms of Natural Rubber Pyrolysis Gas Generation Using RMD Simulations and TG-FTIR Experiments. Energy Convers. Manag. 2019, 189, 143–152. [Google Scholar] [CrossRef]
  52. Vatani, A.; Mehrpooya, M.; Gharagheizi, F. Prediction of Standard Enthalpy of Formation by a QSPR Model. Int. J. Mol. Sci. 2007, 8, 407–432. [Google Scholar] [CrossRef] [Green Version]
  53. Parks, G.S.; Mosher, H.P. Enthalpy and Free Energy Changes in Some Simple Polymerization Processes. J. Polym. Sci. Part A Gen. Pap. 1963, 1, 1979–1984. [Google Scholar] [CrossRef]
  54. Salomatov, V.V.; Kuznetsov, G.V.; Syrodoy, S.V. Influence of the Degree of Coal Metamorphism on Characteristics and Conditions of Ignition of Coal-Water Fuel Drops. Thermophys. Aeromechanics 2018, 25, 773–788. [Google Scholar] [CrossRef]
  55. Arena, U. Process and Technological Aspects of Municipal Solid Waste Gasification. A Review. Waste Manag. 2012, 32, 625–639. [Google Scholar] [CrossRef] [PubMed]
  56. Guan, Y.; Luo, S.; Liu, S.; Xiao, B.; Cai, L. Steam Catalytic Gasification of Municipal Solid Waste for Producing Tar-Free Fuel Gas. Int. J. Hydrogen Energy 2009, 34, 9341–9346. [Google Scholar] [CrossRef]
  57. Donskoy, I.G. Numerical Study on the Efficiency of Biomass and Municipal Waste Fixed-Bed Co-Gasification. E3S Web Conf. 2019, 114, 06006. [Google Scholar] [CrossRef]
  58. Xie, J.; Zhong, W.; Jin, B.; Shao, Y.; Huang, Y. Eulerian–Lagrangian Method for Three-Dimensional Simulation of Fluidized Bed Coal Gasification. Adv. Powder Technol. 2013, 24, 382–392. [Google Scholar] [CrossRef]
  59. Chen, C.J.; Hung, C.I.; Chen, W.H. Numerical Investigation on Performance of Coal Gasification under Various Injection Patterns in an Entrained Flow Gasifier. Appl. Energy 2012, 100, 218–228. [Google Scholar] [CrossRef]
  60. He, C.; Feng, X.; Chu, K.H.; Li, A.; Liu, Y. Industrial-Scale Fixed-Bed Coal Gasification: Modeling, Simulation and Thermodynamic Analysis. Chinese J. Chem. Eng. 2014, 22, 522–530. [Google Scholar] [CrossRef]
  61. Haseli, Y.; van Oijen, J.A.; de Goey, L.P.H. A Detailed One-Dimensional Model of Combustion of a Woody Biomass Particle. Bioresour. Technol. 2011, 102, 9772–9782. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, Q.; Dor, L.; Fenigshtein, D.; Yang, W.; Blasiak, W. Gasification of Municipal Solid Waste in the Plasma Gasification Melting Process. Appl. Energy 2012, 90, 106–112. [Google Scholar] [CrossRef]
  63. Uskov, S.I.; Enikeeva, L.V.; Potemkin, D.I.; Belyaev, V.D.; Snytnikov, P.V.; Gubaidullin, I.M.; Kirillov, V.A.; Sobyanin, V.A. Kinetics of Low-Temperature Steam Reforming of Propane in a Methane Excess on a Ni-Based Catalyst. Catal. Ind. 2017, 9, 104–109. [Google Scholar] [CrossRef]
  64. McAllister, S.; Chen, J.-Y.; Fernandez-Pello, A.C. Chemical Kinetics. In Fundamentals of Combustion Processes; Springer: New York, NY, USA, 2011; pp. 49–73. [Google Scholar]
  65. Matthess, N.; Schweich, D.; Martin, B.; Castagna, F. From Light-off Curves to Kinetic Rate Expressions for Three-Way Catalysts. Top. Catal. 2001, 16–17, 119–124. [Google Scholar] [CrossRef]
  66. Westbrook, C.K.; Dryer, F.L. Chemical Kinetic Modeling of Hydrocarbon Combustion. Prog. Energy Combust. Sci. 1984, 10, 1–57. [Google Scholar] [CrossRef]
  67. Mendoza, J.A.; Hwang, S. Tubular Reactor Design for the Oxidative Dehydrogenation of Butene Using Computational Fluid Dynamics (CFD) Modeling. Korean J. Chem. Eng. 2018, 35, 2157–2163. [Google Scholar] [CrossRef]
  68. Cheng, G.; Farmer, R.; Jones, H.; McFarlane, J. Numerical Simulation of the Internal Ballistics of a Hybrid Rocket Motor. In Proceedings of the 32nd Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 10–13 January 1994; American Institute of Aeronautics and Astronautics: Reston, VA, USA, 1994. [Google Scholar]
  69. Tao, F.; Srinivas, S.; Reitz, R.D.; Foster, D.E. Comparison of Three Soot Models Applied to Multi-Dimensional Diesel Combustion Simulations. JSME Int. J. Ser. B 2005, 48, 671–678. [Google Scholar] [CrossRef] [Green Version]
  70. McAllister, S.; Chen, J.-Y.; Fernandez-Pello, A.C. Droplet Evaporation and Combustion. In Fundamentals of Combustion Processes; Springer: New York, NY, USA, 2011; pp. 155–175. [Google Scholar]
  71. Celik, I.B.; Ghia, U.; Roache, P.J.; Freitas, C.J.; Coleman, H.; Raad, P.E. Procedure for Estimation and Reporting of Uncertainty Due to Discretization in CFD Applications. J. Fluids Eng. Trans. ASME 2008, 130, 0780011–0780014. [Google Scholar] [CrossRef]
  72. Fugmann, H.; Schnabel, L.; Frohnapfel, B. Heat Transfer and Pressure Drop Correlations for Laminar Flow in an In-Line and Staggered Array of Circular Cylinders. Numer. Heat Transf. Part A Appl. 2019, 75, 1–20. [Google Scholar] [CrossRef]
  73. Glushkov, D.O.; Strizhak, P.A.; Vershinina, K.Y. Minimum Temperatures for Sustainable Ignition of Coal Water Slurry Containing Petrochemicals. Appl. Therm. Eng. 2016, 96, 534–546. [Google Scholar] [CrossRef]
Figure 1. Schematic of experimental conditions: (a) first condition; (b) second condition.
Figure 1. Schematic of experimental conditions: (a) first condition; (b) second condition.
Applsci 12 12273 g001
Figure 2. Video frames of ignition and combustion of CLF droplets No. 5 (FC 90% + cardboard 10%) at Tg = 873 K [23] under the following conditions: (a) radiant heating (Vg ≈ 0); (b) convective heating (Vg ≈ 3 m/s).
Figure 2. Video frames of ignition and combustion of CLF droplets No. 5 (FC 90% + cardboard 10%) at Tg = 873 K [23] under the following conditions: (a) radiant heating (Vg ≈ 0); (b) convective heating (Vg ≈ 3 m/s).
Applsci 12 12273 g002
Figure 3. A reacting particle in the multiple surface reactions model [42]: Tp—droplet temperature; T∞—ambient gas temperature; Cd,b and Cd,s—concentration of the diffusion-limited species; Ck—concentrations of all other species.
Figure 3. A reacting particle in the multiple surface reactions model [42]: Tp—droplet temperature; T∞—ambient gas temperature; Cd,b and Cd,s—concentration of the diffusion-limited species; Ck—concentrations of all other species.
Applsci 12 12273 g003
Figure 4. Scheme of the solution domain for ignition in the CLF droplet/air system.
Figure 4. Scheme of the solution domain for ignition in the CLF droplet/air system.
Applsci 12 12273 g004
Figure 5. Establishing ignition-delay times by the CLF droplet temperature evolution as illustrated by composition No. 5 (FC 90% + cardboard 10%) at Tg = 973 K, Vg = 0.
Figure 5. Establishing ignition-delay times by the CLF droplet temperature evolution as illustrated by composition No. 5 (FC 90% + cardboard 10%) at Tg = 973 K, Vg = 0.
Applsci 12 12273 g005
Figure 6. Relative ignition-delay times of CLF No. 4 (FC 90% + plastic 10%) vs. devolatilization constant of the combusting particle at different temperatures of the radiant heating source.
Figure 6. Relative ignition-delay times of CLF No. 4 (FC 90% + plastic 10%) vs. devolatilization constant of the combusting particle at different temperatures of the radiant heating source.
Applsci 12 12273 g006
Figure 7. Relative ignition-delay times of CLF No. 4 (FC 90% + plastic 10%) vs. reaction heat fraction absorbed by the solid of the combusting particle at different temperatures of the radiant heating source.
Figure 7. Relative ignition-delay times of CLF No. 4 (FC 90% + plastic 10%) vs. reaction heat fraction absorbed by the solid of the combusting particle at different temperatures of the radiant heating source.
Applsci 12 12273 g007
Figure 8. Relative ignition-delay times of CLF vs. volatile component fraction/combustible fraction ratio of the combusting particle at different temperatures of the radiant heating source.
Figure 8. Relative ignition-delay times of CLF vs. volatile component fraction/combustible fraction ratio of the combusting particle at different temperatures of the radiant heating source.
Applsci 12 12273 g008
Figure 9. Time dependences of maximum temperature of CLF droplets No. 5 (FC 90% + cardboard 10%) under radiant heating (Vg ≈ 0): (a) simulation results at different heating temperatures; (b) simulation results vs. experimental data [41] at Tg = 1273 K.
Figure 9. Time dependences of maximum temperature of CLF droplets No. 5 (FC 90% + cardboard 10%) under radiant heating (Vg ≈ 0): (a) simulation results at different heating temperatures; (b) simulation results vs. experimental data [41] at Tg = 1273 K.
Applsci 12 12273 g009
Figure 10. Ignition-delay times vs. heating source temperature at Vg = 0 (left side) and Vg = 3 m/s (right side) for CLF particles of different compositions: (a) No. 1 (FC 100%); (b) No. 2 (FC 90% + wood 10%); (c) No. 3 (FC 90% + rubber 10%); (d) No. 4 (FC 90% + plastic 10%); (e) No. 5 (FC 90% + cardboard 10%).
Figure 10. Ignition-delay times vs. heating source temperature at Vg = 0 (left side) and Vg = 3 m/s (right side) for CLF particles of different compositions: (a) No. 1 (FC 100%); (b) No. 2 (FC 90% + wood 10%); (c) No. 3 (FC 90% + rubber 10%); (d) No. 4 (FC 90% + plastic 10%); (e) No. 5 (FC 90% + cardboard 10%).
Applsci 12 12273 g010aApplsci 12 12273 g010b
Figure 11. Temperature fields and videoframes obtained in experiments of ignition and combustion of CLF droplets No. 5 (FC 90% + cardboard 10%) at Tg = 873 K [23] under the following conditions: (a) radiant heating (Vg ≈ 0); (b) convective heating (Vg ≈ 3 m/s).
Figure 11. Temperature fields and videoframes obtained in experiments of ignition and combustion of CLF droplets No. 5 (FC 90% + cardboard 10%) at Tg = 873 K [23] under the following conditions: (a) radiant heating (Vg ≈ 0); (b) convective heating (Vg ≈ 3 m/s).
Applsci 12 12273 g011aApplsci 12 12273 g011b
Figure 12. Ignition-delay times vs. heating source temperature for CLF composition No. 5 (FC 90% + cardboard 10%) at: (a) Vg = 0; (b) Vg = 3 m/s.
Figure 12. Ignition-delay times vs. heating source temperature for CLF composition No. 5 (FC 90% + cardboard 10%) at: (a) Vg = 0; (b) Vg = 3 m/s.
Applsci 12 12273 g012
Table 1. Characteristics and properties of composite liquid fuel components [28].
Table 1. Characteristics and properties of composite liquid fuel components [28].
ComponentWa
(%)
Ad
(%)
Vdaf
(%)
Qas,V
(MJ/kg)
Specific Heat Capacity (kJ/(kg∙K))Thermal Conductivity (W/(m∙K))
Filter cake C (dry)-26.5023.1024.831.300.366
Wood20.002.0083.1016.451.550.200
Rubber2.001.8067.4033.501.380.150
Plastic2.000.2099.5022.002.300.260
Cardboard5.003.0096.5717.502.300.130
Table 2. Ultimate analysis of composite liquid fuel components [28].
Table 2. Ultimate analysis of composite liquid fuel components [28].
ComponentCdaf (%)Hdaf (%)Ndaf (%)Sdaf (%)Odaf (%)
Filter cake C (dry)87.205.102.101.104.50
Wood50.306.000.200.1043.40
Rubber97.901.200.300.60-
Plastic66.707.90--25.40
Cardboard46.306.300.300.2046.90
Table 3. Average ignition-delay times of CLF exposed to radiant heating (Vg ≈ 0) at different air temperatures [23].
Table 3. Average ignition-delay times of CLF exposed to radiant heating (Vg ≈ 0) at different air temperatures [23].
No. of
Composition
Temperature of Motionless Air (Vg ≈ 0)
723 K773 K873 K973 K1073 K1173 K1273 K
No. 119.5716.3211.398.145.904.132.98
No. 218.1515.0410.477.505.263.772.69
No. 317.0014.089.676.744.653.292.31
No. 416.5513.609.196.264.252.982.02
No. 514.1511.357.304.622.851.730.96
Table 4. Average ignition-delay times of CLF exposed to convective heating (Vg ≈ 3 m/s) at different air temperatures [23].
Table 4. Average ignition-delay times of CLF exposed to convective heating (Vg ≈ 3 m/s) at different air temperatures [23].
No. of
Composition
Air-Flow Temperature (Vg ≈ 3 m/s)
723 K773 K823 K873 K923 K973 K
No. 17.486.004.863.933.142.56
No. 27.035.624.533.652.892.36
No. 36.655.384.383.532.812.31
No. 46.545.284.263.432.712.19
No. 56.375.154.143.342.652.14
Table 5. Kinetic parameters of chemical reactions of CLF ignition.
Table 5. Kinetic parameters of chemical reactions of CLF ignition.
ReactionArEr (kJ/(mol·K))Ref.
Water Evaporation
H2O(l) = H2O(g)5.13∙10687.9[49,50]
Pyrolysis
Biomass pyrolysis
lignin = 0.209CO2 + 0.396CO + 0.109H2 + 0.249H2O + 0.037vol2.202∙1012181[50]
hemicellulose = 0.209CO2 + 0.396CO+0.109H2 + 0.249H2O + 0.037vol2.527∙1011147[50]
cellulose = 0.209CO2 + 0.396CO + 0.109H2 + 0.249H2O + 0.037vol1.379∙1014193[50]
Rubber pyrolysis
rubber = 0.8H2 + 0.0009C2H4 + 0.194CH4 + 0.0025C3H6 + 0.0018C4H6 + 0.0008C2H25.5∙1018181[51,52]
Plastic pyrolysis
polyethylene = 0.825H2 + 0.07C2H4 + 0.05CH4 + 0.03C3H6 + 0.02C2H6 + 0.005C3H815∙10340[53]
Gasification
C + H2O = CO + H22.07∙107220[50,54,55,56]
C + CO2 = 2CO1.32∙107259[50,55,56,57,58,59]
C + 2H2 = CH45∙10630[50,55,56,58]
C + 2H2O = CO2 + 2H22.1∙106158[50,57,60]
CO2 + H2 = CO + H2O5∙10630[49,58,61]
Combustion
C + O2 = CO22∙101260.6[55,57,62]
C + 1/2O2 = CO2∙101260.6[55,57,58,59]
H2 + 1/2O2 = H2O2.1∙1014129.8[54,55,58]
2CO + O2 = 2CO21.4∙101396.8[54,55,58]
CO + H2O = H2 + CO25∙10630[55,56,58]
CO + 3H2 = CH4 + H2O5∙10630[49]
C + H2O = 1/2CO2 + 1/2CH45.6∙101236.2[49,63]
CH4 + 2O2 = CO2 + 2H2O5.6∙1012103.8[58,64]
C2H4 + O2 = 2CO + 2H21∙1012173[21]
2C3H6 + 9O2 = 6CO2 + 6H2O1.51∙101585.6[65]
2C2H6 + 7O2 = 4CO2 + 6H2O1.1∙1012125.52[66]
2C3H8 + 10O2 = 6CO2 + 8H2O8.6∙1011125.52[66]
C4H6 + 3O2 = 4CO + 2H2O + H28.8∙1011126.37[67,68]
C4H8 + 1/2O2 = C2H4 + H2O6∙1012502[67]
C4H8 + 6O2 = 4CO2 + 4H2O3∙10710.4[67]
C4H6 + 11/2O2 = 4CO2 + 3H2O3∙10710.4[67]
C2H2 + O2 = 2CO + H26∙101350[69]
Table 6. The main characteristics of different CLF compositions.
Table 6. The main characteristics of different CLF compositions.
CharacteristicsNo. of Composition
No. 1No. 2No. 3No. 4No. 5
Q (MJ/kg)10.7811.2913.0511.8811.43
Specific heat capacity (kJ/(kg∙K))2.46002.41752.35492.41362.4136
Thermal conductivity (W/(m∙K))0.47960.45820.45500.46200.4547
Viscosity (Pa∙s)0.5570.5570.5570.5570.557
Wa (%)40.0038.0036.2036.2036.50
Ad (%)26.5024.0524.0323.8724.15
Vdaf (%)23.1029.1027.5330.7430.45
Cdaf (%)87.2083.5188.2785.1583.11
Hdaf (%)5.105.194.715.385.22
Ndaf (%)2.101.911.921.891.92
Sdaf (%)1.101.001.050.991.01
Odaf (%)4.508.394.056.598.74
Table 7. Grid convergence analysis.
Table 7. Grid convergence analysis.
ParametersGrid 1Grid 2 (Examined)Grid 3
Number of cells562514,400360,000
Element size4·10−42.5·10−40.5·10−4
Ignition-delay time of composite fuel No. 5 (at 1173 K, Vg = 0), s1.9951.9901.931
GCI (%)0.0620.0360.001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Antonov, D.; Glushkov, D.; Paushkina, K.; Kuznechenkova, D.; Ramanathan, A. A Mathematical Model of Industrial Waste-Derived Fuel Droplet Combustion in High-Temperature Air. Appl. Sci. 2022, 12, 12273. https://doi.org/10.3390/app122312273

AMA Style

Antonov D, Glushkov D, Paushkina K, Kuznechenkova D, Ramanathan A. A Mathematical Model of Industrial Waste-Derived Fuel Droplet Combustion in High-Temperature Air. Applied Sciences. 2022; 12(23):12273. https://doi.org/10.3390/app122312273

Chicago/Turabian Style

Antonov, Dmitrii, Dmitrii Glushkov, Kristina Paushkina, Daria Kuznechenkova, and Anand Ramanathan. 2022. "A Mathematical Model of Industrial Waste-Derived Fuel Droplet Combustion in High-Temperature Air" Applied Sciences 12, no. 23: 12273. https://doi.org/10.3390/app122312273

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