Next Article in Journal
Experimental and Statistical Analysis of Concrete Eco-Cobble Using Organic and Synthetic Fibers
Previous Article in Journal
Experimental Study on the Suspending Mechanism of Suspending Agent in Coal-Based Solid Waste Slurry for Long-Distance Pipeline Transportation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Energy Evaluation and Mathematical Modeling of Pellet Production from Metal-Bearing Waste with a Focus on Alternative Applications of Reducing Agents

1
Institute of Metallurgy, Faculty of Materials, Metallurgy and Recycling, Technical University of Kosice, Letna 9, 042 00 Kosice, Slovakia
2
Department of Process Technique, Faculty of Manufacturing Technologies with a Seat in Presov, Technical University of Kosice, Bayerova 1, 080 01 Presov, Slovakia
3
Faculty of Engineering Materials, Silesian University of Technology, Krasińskiego 8, 40-019 Katowice, Poland
*
Author to whom correspondence should be addressed.
Processes 2024, 12(9), 1938; https://doi.org/10.3390/pr12091938
Submission received: 9 August 2024 / Revised: 31 August 2024 / Accepted: 6 September 2024 / Published: 9 September 2024
(This article belongs to the Special Issue Pyrolytic Process for Recycling)

Abstract

:
The authors of this study focused on the energy and material assessment of processes for processing pellets from metal-bearing waste, specifically Fe concentrate. A mathematical model was created for process evaluation, with which thermotechnical calculations of parameters in the processing of metallized pellets were carried out. Thermodynamic calculations were performed to determine the enthalpy of the charge in individual devices (drying chamber, rotary kiln, cooler). For the reduction of Fe oxides, carbon from coke (with Fe oxide reductions of 50%, 61%, and 92%) and lignite (with Fe oxide reductions of 69% and 92%) were considered as part of the pellets. The degree of reduction of iron oxides was a determining parameter, and the consumption of the reducing agent corresponded to the direct reduction of Fe oxides by carbon with a coefficient of 1.5. Another determining parameter was the input and output temperature in individual devices. For a more precise description of the processes in individual devices, calculations were carried out zonally. The results of the calculations are analyses and recommendations for feasible alternatives for the reducing agent and associated processes.

1. Introduction

Natural resources are a valuable commodity in every country, including those outside the European Union. In the past, heavy industry and raw material extraction prevailed in Slovakia. The mining industry, in particular, gained significant prominence in the national economy due to increasing productivity. However, since 2000, there have been signs of a decline resulting from national and local conditions in the mining industry [1]. More recently, these sectors have been phased out, but considerable environmental burdens remain, which have not been addressed for a long time.
The use of resources, production, consumption, and waste have become key topics in circular economy strategies and relevant policies in recent years. These initiatives aim to close the material cycle to maintain the value of products, materials, and resources in the economy for as long as possible. This effectively reduces waste production and the consumption of primary raw materials, leading to decreased environmental pressures [2].
The extraction, production, use, and disposal of resources, as well as waste production and processing, create significant environmental pressure. Environmental policy aims to reduce material consumption in the economy, use resources more efficiently, reduce waste production, and turn waste into a resource [2].
Waste in the European economy is increasingly valued as a valuable resource that can be further utilized. The increase in waste recycling gradually reduces the amount of waste deposited in landfills. Nevertheless, there are considerable differences in waste management among individual countries [2].
European environmental policy, within the framework of the Environmental Action Programme (EAP), has long set goals, including increasing resource efficiency. Priority targets have been set for 2030 and the necessary conditions for their achievement [3].
In line with environmental policy, it is in every state’s interest to reduce the amount of waste that can be further utilized. Large companies, such as metallurgical complexes, are currently analyzing individual production processes and seeking environmental solutions to minimize or eliminate the waste generated.
The production of pellets from metal-bearing materials is an energy-intensive and environmentally demanding process, and research in this area remains relevant. To optimize pellet production, parameters affecting individual processes were studied either through experimental measurements or by creating mathematical models. The properties of pellets processed in thermal aggregates, such as tubular rotary kilns, must meet certain conditions, including granulation, moisture, strength of raw pellets, resistance to rapid heating, and reducibility of oxides in the pellets [4,5].
In preparing raw pellets, it is important to know the properties of metal-bearing materials based on metallographic analyses of samples, which are the basis for the preparation of raw pellets. Laubertová et al. [6] emphasized that the quality of the material depends on the sampling method, which also influences the result of the chemical analysis. This is particularly important when using waste materials that exhibit a certain heterogeneity, thereby affecting the final parameters of the raw pellets. A significant parameter is the strength of the raw pellets, which can be influenced by adding reducing agents, additives, or colloidal particles that harden upon drying, forming mortar-like bridges that increase pellet strength. Otherwise, it is necessary to increase the addition of binding agents such as bentonite or cement and other [5].
Mathematical modeling of processes has focused mainly on the thermal processes that occur during processing. The authors [7] created a simulation model of temperature distribution in pellets and flue gases based on conduction, convection, and radiation in a tubular rotary kiln. Larsson [8] created a model of flow in a rotary kiln and the exit of the flame and flue gases from the burner, addressing the aerodynamics of flow and fuel combustion using CFD models. The heat transfer from the flame in the rotary kiln was also addressed by authors [9] who created a one-dimensional model capable of predicting axial temperature profiles of the flame and wall, and axial profiles of heat flux to the bed of solids and the refractory wall. Interesting models are also described in work [10] on the oxidation of magnetite during the hardening of iron ore pellets, and [11], which address the physicochemical processes occurring during the hardening of pellets. Despite the mathematical models, experimental research in the field of unconventional technologies is also important, as presented by the authors [12] in the preparation of Carbofer-type pellets. Among the interesting unconventional studies is the research published in [13], which focused on the use of plastics as a reducing agent. Using plastics, a degree of metallization greater than 90% was achieved with reduction in 30 min. The created mixture exhibited better reactivity, and syngas was simultaneously produced during thermal processing.
In the realm of environmental technologies, the application of hydrogen as a significant carbon-free reductant is being explored during the thermal processing of raw pellets and the reduction of oxides in the pellets. The goal is to reduce the production of CO2 greenhouse gases in these technologies. Experimental research using hydrogen was presented in a research report [14], where hydrogen was used in combination with coke, resulting in a high degree of iron oxide reduction with 3.1 wt.% coke. In the study [15], the authors focused on using pure hydrogen to reduce pellets, and the research results indicate further possibilities for investigation in this area. Numerous other publications demonstrate that research in this field remains a highly discussed topic, with new opportunities emerging within environmental policy frameworks.
The authors aimed to investigate the energy assessment of pellet production from leach residue from Sereď, Slovakia. Leach residue is essentially an Fe concentrate suitable for the production of pellets for use in iron metallurgy. Research on the utilization of leach residue has been conducted previously, with results published in [16,17,18]. A subsequent study [14] evaluated the possibilities of pellet production and properties based on supplied samples. The results of the pre-reduction of pellets in a tubular rotary kiln were published in [19]. Despite positive outcomes, no project for utilizing the waste leach residue from the resulting waste heap was implemented.
However, renewed interest from investors led to revisiting the potential realization of this project at a decommissioned metallurgical plant, considering the revival of production. To support the implementation of this process, a mathematical model was developed for thermotechnical calculations involved in processing leach residue into raw pellets. The goal of the production process is to increase the metallic Fe content in the pellets through direct reduction of Fe oxides.
This paper presents a detailed approach to the material and energy evaluation of this process, focusing on the potential revival of production at the decommissioned plant. The authors conducted theoretical research to develop a mathematical model for processing leach residue. The Introduction provides background on the research and the mathematical methods used to model the physicochemical processes during pellet production.
In the Materials and Methods Section, the chemical composition of ore processed at NH Sereď is discussed, with laboratory tests showing that increasing coke content enhances Fe reduction efficiency. The leach residue sample, labeled Fe concentrate SE52, was analyzed and used as input data for the mathematical model. Tests indicated that the optimal amount of solid reducing agent is 1.5 times the theoretical carbon requirement.
This paper also describes the equipment required for processing raw pellets, including a technological scheme for pellet production in Nižná Slaná, Slovakia. Material and heat balances were calculated for various stages, such as drying, preheating, firing in the rotary kiln, and cooling, with thermodynamic calculations determining the enthalpy of the charge in each device. A detailed description of the processes and the main physical equations used in the model is provided.
The results, presented in tables, show the composition of raw and processed pellets and reveal that carbon was not completely consumed during reduction. Several physical parameters were calculated for different processing strategies, showing that producing pellets with a reduction degree of Fe oxides between 50–65% is feasible. The rotary kiln parameters were found suitable for processing raw pellets from the Fe concentrate in Sereď leach residue, supporting the potential realization of the project.

2. Materials and Methods

This study of raw pellet processing was based on the composition of waste after processing nickel ore at NH Sereď. NH Sereď processed ore from Albania, which contained 1.05% Ni, 0.06% Co, 49% Fe, and 2.6% Cr. The production was primarily focused on the extraction of Ni and Co, which were present in low concentrations. Despite the unprofitable production, there was a requirement to process the ore. Due to societal changes and the unprofitability of production, the facility was closed in 1992. Before its closure, NH Sereď produced 2500 tons of Ni and 60 tons of Co annually through hydrometallurgical processing. The process also resulted in a leach residue heap, following ammoniacal leaching, with high iron and chromium content, weighing approximately 6 million tons [20].
The resulting waste was stored in heaps, and due to its properties, it was very difficult to reclaim. The waste storage also posed a significant environmental burden, particularly due to its migration into the surrounding area caused by wind action. For this reason, a deeper analysis of this waste was undertaken, and the results indicate that it is suitable for further processing.
Studies on the properties of metal-bearing waste from the heap focused on its technological feasibility for processing. Technologies for thermal treatment were considered, including agglomeration on a sintering belt and its use in the form of pellets. The agglomeration process could be carried out directly at the metallurgical plant, where iron ore is primarily processed. Based on the results of the studies [16,17,18,20] and subsequent pilot-scale tests [19], the decision was made to utilize pelletization technology and the subsequent thermal treatment of raw pellets. Laboratory tests demonstrated that the input material had suitable parameters for producing raw pellets with sufficient strength and resistance to rapid heating without cracking during heating.

2.1. Leach Residue—Metal-Bearing Waste

The leach residue obtained as a byproduct of the hydrometallurgical processing of Ni ore is an inherently metal-bearing material with high Fe content, making it suitable for further processing in the form of pellets. Laboratory tests conducted on collected samples demonstrated the potential for its further processing.
Based on the thermodynamic analysis and laboratory tests of Fe concentrate reduction from Sereď, it is evident that the direct reduction of iron oxides using a solid internal reducing agent is the most advantageous method.
Based on the experimental tests with leach residue from Sereď mentioned in [14], it can be concluded that a 60% reduction of Fe is achieved in approximately 60 min. From Figure 1, it is evident that the direct reduction of Fe from Fe concentrate using a solid reducing agent—coke dust present in the pellets—achieved high efficiency of 81.02%, provided that the pellets contained a sufficient amount of reducing agent and the reduction process lasted for an adequate duration of 180 min [14].
Figure 1 illustrates the results of laboratory tests that monitored Fe reduction efficiency using coke in the pellets and in combination with hydrogen. The obtained results indicate that increasing the coke content leads to higher Fe reduction efficiency. The dependencies show a logarithmic relationship, which is also characteristic when combined with hydrogen reduction.
The calculation of average values for further computations was based on the composition of dry leach residue. The calculations account for 12.322% moisture content relative to 100% dry leach residue.
Based on the sieve analysis, the granulometry of the leach residue was determined as follows: greater than 9.53 mm/0%; greater than 6.35 mm/0%; greater than 0.149 mm/8.5%; less than 0.149 mm/91.5%.
For the preparation of this study, a sample was taken from the leach residue heap in Sereď, whose composition is shown in Table 1. The sample was labelled as Fe concentrate SE52, and average values from the minimum and maximum values were used for the calculations, serving as input data in the mathematical model. Laboratory tests revealed that the optimal amount of solid reducing agent is 1.5 times the theoretical amount of carbon needed to reduce the iron oxides present in the Fe concentrate.
The calculations considered solid reducing agents coke and lignite. These agents were selected based on their immediate availability, either produced or mined in Slovakia. The average compositions of the reducing agents are shown in Table 2.
From the composition of the reducing agents in Table 2, it is possible to compare the carbon content for direct reduction, which is 28.59 wt.% lower in lignite, thereby affecting the extent of direct reduction of oxides. The direct reduction of iron oxides from the Fe concentrate begins at a temperature of 956 °C, but according to laboratory tests, the optimal reduction temperature is considered to be between 1000 and 1100 °C.
Within this temperature range, it is expected that the gaseous product of the direct reduction of iron oxides will mainly be CO, which will need to be further oxidized to carbon dioxide outside the furnace’s working chamber. The generated CO will simultaneously create a reducing atmosphere in the rotary kiln, which will support the reduction of Fe oxides while preventing reoxidation.
Due to the high content of gangue components, such as CaO, SiO2, and Al2O3, in the Fe concentrate from Sereď, the metal content of the reduced agglomerate or pellets can be increased above the theoretical limit concentration of 68% Fe only by adding iron-bearing waste, such as mill scale or converter dust or sludge, to the Fe concentrate. This method can increase the iron content in the reduced agglomerate or pellets to up to 86% Fe.
Due to the high temperatures, complex chemical and physical processes, as well as polymorphic transformations, occur in the calcined raw pellets. These processes determine the mechanical and metallurgical properties of the fired and cooled pellets. Since these processes reduce the particle surface area and the free surface energy of the system, sintering is an irreversible process [4].
During firing, changes occur in both the metal-bearing and gangue parts of the charge:
(a)
In the metal-bearing part of the charge, magnetite may oxidize to hematite, and recrystallization of magnetite and hematite may occur;
(b)
The oxides in the gangue part of the concentrate react with each other at lower temperatures in the solid state and during the transition from solid to liquid state, forming slag;
(c)
Additionally, reactions occur in both the solid and liquid states between the oxides in the metal-bearing and gangue parts of the concentrate.

2.2. Processes in Technological Equipment

The process of processing raw pellets requires the use of technology that includes various primary and support equipment. The most important equipment is the thermal aggregate—a rotary kiln with a burner system to provide thermal input for their processing. The rotary kiln is the basis of the mathematical model for further heat technical calculations in the other equipment.
For the proposed process of processing raw pellets from Fe concentrate, the thermal units in Nižná Slaná, Slovakia, are planned to be utilized. The technological operation in Nižná Slaná was shut down and is currently being considered for use in processing metal-bearing waste from metallurgical operations. Consideration is also given to their modification and the addition of auxiliary equipment for the processing of metal-bearing waste. This technology served as the basis for developing a mathematical model for the material and energy balance of processing raw pellets from metal-bearing waste, such as the described Fe concentrate from the waste heap after nickel ore processing.
Figure 2 shows the existing condition of the decommissioned technology, highlighting the placement of the burner system and the discharge of pellets from the rotary kiln into the cooler. As observed in Figure 2, the technology requires certain investments, but it can be brought back into operation after modifications. For this reason, an evaluation of the technology was undertaken in terms of material and energy aspects for processing waste leach residue from the heap in Sereď.
The burner system can be adapted for both gaseous and liquid fuels. Due to the short length of the rotary kiln, it is more advantageous to use gaseous fuel, also because of its availability from the distribution network. The calculations used the composition of natural gas as provided in [21], with the adjusted composition listed in Table 3.
In the calculations of individual aggregates, from drying and preheating of the charge through firing in the rotary kiln to cooling, material and heat balances were used in these devices in combination with the kinetics of heat transfer within each device. Within the heat balance, thermodynamic calculations were carried out to determine the enthalpy of the charge in each device. In the process of reducing Fe oxides, carbon from coke and lignite, which are part of the charge, is considered. A mathematical model was developed in Excel to determine all the necessary thermotechnical parameters.
For more economical use of flue gas energy generated in the process of processing metal-bearing waste, a cogeneration unit was proposed and incorporated into the technological scheme of the process in some alternatives.
Using the created mathematical model, calculations of the following pellet production alternatives were carried out:
  • Reducing agent—coke: reduction of Fe oxides in pellets to 50% (C_50R).
  • Reducing agent—coke: reduction of Fe oxides in pellets to 61% (C_61R).
  • Reducing agent—coke: reduction of Fe oxides in pellets to 92% (C_92R).
  • Reducing agent—lignite: reduction of Fe oxides in pellets to 69% (BC_69R).
  • Reducing agent—lignite: reduction of Fe oxides in pellets to 92% (BC_92R).
  • Production of oxide pellets without reduction of Fe oxides (OX).
Figure 3 shows the basic technological process scheme for producing pellets from raw pellets. The processing of raw pellets begins in the drying chamber, where the raw pellets are heated and moisture is removed by controlling the temperature in the individual zones Z1, Z2, and Z3. Flue gases from the rotary kiln are introduced into these zones, and after separating the dust particles in the cyclone, they proceed to the afterburning chamber (DSK), where unburned gaseous components CO and H2 are combusted along with gaseous fuel and combustion air. The flue gases from DSK continue to the mixer (M), where the temperature of the flue gases is adjusted to the required level before entering Z3. In the distributor after M, part of the mixed flue gases is directed into Z3, and another part continues to another mixer (M), where the temperature is further adjusted for Z2. The flue gases from Z3 and Z2, after passing through the raw pellets, are collectively led to the mixer (M), where their temperature is reduced by mixing with air to the required level before entering Z1 in the drying chamber. The flue gases from Z1, after passing through the raw pellets, are directed to the filter (F1) for cleaning before entering the chimney (K), where the dust particles from the pellets in the drying chamber are captured.
After the raw pellets are dried and heated in the drying chamber, they proceed to thermal processing in the rotary kiln, where their temperature is increased, and the actual induration process occurs, along with the reduction of iron oxides using reducing agents. The flue gases and the reduction process itself create a reducing atmosphere, which also influences the reduction or metallization process of the pellets. Gaseous fuel and a portion of preheated air from the cooler are used as heat sources.
In the cooler, the pellets from the rotary kiln, which are at a high temperature, are cooled to the desired temperature by introducing air from the surroundings. The hot air from the cooler is distributed to the rotary kiln and DSK for combustion.
Figure 4 presents a modified technological scheme, which is an enhanced version of the scheme in Figure 3, with the addition of a cogeneration unit (KG) after DSK and M. The modification of the scheme was proposed based on calculations of the required amounts of flue gases and air for mixing into the heating zones. The calculations revealed that in the case of a large volume of flue gases, it is not possible to reduce the temperature simply by mixing in air in M, but it is necessary to remove the heat by other means. To utilize the waste heat more efficiently, a KG unit without internal combustion was proposed. In this case, it is necessary to identify a type of heat exchanger that will not get clogged by dust particles, or at least one that is easy to maintain. A similar issue was discussed by the authors in [22], where they addressed the design of a heat exchanger for thermal air engines and their use in combined heat and power production. An ORC was also considered for the KG.

Brief Description of the Processes Occurring in Individual Equipment

  • Drying chamber
Basic processes occurring during the preheating of raw pellets:
Z1—Preheating I
Heating raw pellets from ambient temperature to an estimated temperature of approximately 110–130 °C.
Evaporation of water from the raw pellets.
Decrease in the mass of raw pellets due to the evaporated water.
Increase in the volume of flue gases due to the evaporated water and solid particles from the charge passing through the preheating zone.
Z2—Preheating II
Heating and possibly further drying the raw pellets to approximately 230 °C.
The mass of the raw pellets does not change.
The volume of flue gases does not change.
Z3—Preheating III
Heating raw pellets from 230 °C to approximately 400–650 °C.
The mass of raw pellets decreases due to the combustion of sulfur in the raw pellets.
The volume of flue gases may change due to the combustion of sulfur and possibly some volatile substances.
  • Rotary kiln
Basic processes occurring in the rotary kiln:
The charge is heated from 450–650 °C to 1100 °C.
As a result of the reduction of Fe oxides, CO is released into the flue gases, which partially burns to CO2 in a slightly oxidizing atmosphere.
There is partial combustion of excess carbon from the raw pellets to CO2 or CO.
The reduced charge exits to the cooler.
The primary source of thermal energy is supplied by burners (natural gas with excess combustion air, utilizing air from the charge cooler, preheated to approximately 400 °C).
The temperature of the flue gases at the kiln outlet varies depending on the specific alternative.
The flue gases at the kiln outlet, in addition to the components of complete fuel combustion, will contain the following:
  • Unburned CO from the reduction of oxides,
  • Solid particles from the raw pellets,
  • Due to the reduction of Fe oxides, the combustion of part of the excess carbon from the raw pellets, and additional solid particles, the mass of the charge decreases.
  • The resulting amount of flue gas is further cleaned, afterburned, and possibly mixed with air for the process of preheating and drying the charge.
  • Cooler—pellet cooling
Basic processes occurring in the charge cooler:
The charge is cooled from approximately 1100 °C to 410 °C.
The primary cooling medium is air, which is heated from ambient temperature to approximately 380 °C.
Additional cooling can be achieved by using water to cool part of the shell.
  • The mass of the pellets slightly decreases due to solid particles from the charge.
  • The preheated air is used as combustion air in the rotary kiln or the combustion chamber.
  • Combustion/Afterburner chamber DSK
Basic processes occurring in the DSK:
In the combustion chamber, the flammable components of the flue gases (CO, etc.) are fully combusted, raising the temperature of the flue gases. In the case of oxide pellets, natural gas is burned to provide the necessary amount of heat for the preheating and drying of raw pellets.
The source of thermal energy is supplied by burners (natural gas), with complete combustion of these fuels and the flammable components of the flue gases (CO, etc.) assumed, with the appropriate excess of combustion air. Air from the charge cooler or air from the surrounding atmosphere is used for combustion.
  • The quantity and composition of the flue gases change;
  • The flue gases at the outlet contain only components of complete combustion;
  • The flue gases at the exit temperature proceed to Mixer I for cooling.
  • Mixer I
Basic processes occurring in Mixer M after the DSK:
The flue gases from the DSK are cooled to approximately 800 °C;
The cooling medium is air at ambient temperature.
  • The quantity and composition of the flue gases change.
  • Cogeneration/Heat exchanger
Basic processes occurring during cogeneration:
If necessary, the flue gases from Mixer I are cooled to approximately 500 °C before entering Z3 preheating in the case of a reducing agent—lignite.
Part of the thermal energy from the flue gases is used to generate electricity in the cogeneration unit (KG).
  • The quantity and composition of the flue gases change.
  • Mixer II
Basic processes occurring in Mixer M:
A portion of the flue gases from Mixer I is cooled to approximately 400 °C before entering Z2.
The cooling medium is air at ambient temperature.
  • The quantity and composition of the flue gases change.
  • Mixer III
Basic processes occurring in Mixer M before entering Z1:
The flue gases, after passing through Z2 and Z3, are mixed and cooled to approximately 300 °C.
The cooling medium is air at ambient temperature.
  • The quantity and composition of the flue gases change.

2.3. Calculation of the Energy Requirements of the Processes

The energy requirements of the processes involved in processing raw metallic pellets also determine the economic aspect of these processes. Describing the processes necessary or characterizing the needs related to the overall processing of metallic pellets can be approached as partial processes. The outputs of these partial processes are important control points for the subsequent energy processes.
In calculating the energy requirements, a mathematical model of the pellet production process in a rotary kiln was developed. This model covers the entire process of processing raw pellets from Sereď residue, including firing and cooling.
The mathematical model is based on the analysis of technology, material and energy balances, and heat flows of the processes occurring in the individual devices during the production of pellets from raw pellets made from Sereď residue.
The developed mathematical model incorporated equations and modules that have either been previously published or were formulated using the necessary literature sources. The fundamental relationship for determining the thermal content of gases is the temperature dependence of the calculation of the specific heat capacity of flue gases and air, which is given by Equation (1):
c p = a 0 + a 1 T + a 2 T 2 + a 3 T 2 ;       ( k J / ( m 3 K ) )
The coefficients a 0 , a 1 , a 2 , and a 3 in Equation (1) were derived based on the regression of tabulated values of the mean specific heat capacity and are published in [23,24].
The primary calculation was the enthalpy calculation of the heat of pellets, which was based on the composition of raw pellets and the reactions that occurred according to Hess’s law [25]:
Δ H = Δ H p r o d u c t s Δ H r e a c t a n t s
Δ H r e a c t a n t s —enthalpy of reactants input to the process;
Δ H p r o d u c t s —enthalpy of products output from the process.
A temperature dependence calculation model was used for the calculation:
Δ H T 0 = Δ H 298 0 + Δ a T + Δ b 2 T 2 Δ c T + Δ d ;       ( J / m o l )
The coefficients a, b, c, and d are coefficients from [25].
The enthalpy balance of heat during the thermal processing of pellets was adjusted and supplemented by an additional term, and then the change in enthalpy was calculated according to the following relationship:
Δ H T 0 = Δ H 298 0 + Δ a T + Δ b 2 T 2 Δ c T + Δ d Δ H p h a s e 0 ;       ( J / m o l )
For the combustion process and the composition of flue gases, a combustion statics model published in [24] was used. This model allows for the calculation of the composition of the resulting flue gases during the combustion of gaseous fuel, as well as the combustion temperature, which is necessary for calculations in other heat transfer modules.
The composition of flue gases and resulting gases in the individual devices was calculated based on combustion equations and reduction processes from the pellets. It is assumed that CO2 is released from the reducing agent by the combustion of carbon, H2O from the reaction with hydrogen, and SO2 from the sulfur in the pellets. The composition of the flue gases also changed due to dilution with air when lowering the flue gas temperature to the required technological temperature.
Mathematical relationships for convection, conduction, and radiation, as published in [23,26,27,28], were used for the heat transfer calculations. Modules for individual heat transfer processes were developed. The module for heat transfer by radiation was published in [29], which also includes a convection calculation model for industrial furnaces. As part of the thermal calculations, the heat loss to the surroundings from the surface of equipment is calculated as thermal loss, based on the geometry and the overall heat transfer coefficient for radiation and convection.
Radiative heat transfer
Radiative heat transfer constitutes a significant portion of direct heat transfer in furnaces. As mentioned in [29] and supported by experimental data, convective heat transfer accounts for only a quarter of the heat transferred to the charge, with the remainder occurring through radiation. This transfer is particularly significant at higher temperatures.
In radiative heat transfer, it is important to distinguish between the radiation from solid bodies, such as furnace walls, and the radiation from gases.
(a) Gas Radiation
When developing a model for gas radiation, the internal dimensions of the equipment and the associated processes were considered. Experimental measurements revealed that gases with three or more atoms play the most significant role in gas radiation. In combustion, this primarily includes CO2, H2O, and SO2. More information can be found in the cited literature [23,26,30,31].
Because gases emit and absorb heat energy throughout their entire volume, their ability to do so depends not only on temperature but also on the number of molecules that the heat ray encounters along its path. The number of molecules is determined by the thickness of the gas layer and the partial pressure of its components (CO2, H2O, SO2, etc.). Thus, the emissivity or luminosity of the gas is a function of temperature and the product of the partial pressure and the thickness of the gas layer.
ε g = f T g , p p . l v p
Tg—absolute gas temperature, (K);
pp—partial gas pressure, (Pa);
lvp—thickness of the gas layer, (m).
The thickness of the radiating layer in different directions of the workspace is different. Therefore, for technical calculations, the term effective beam length was introduced, which is determined according to the relationship [23,26]:
l e f = η . 4 . V F V
η —correction coefficient, (0.8 ≈ 0.9);
V —volume filled with glowing gas (flue gases), (m3);
F V —the area of the walls delimiting this space, (m2).
Various mathematical models have been developed to determine the emissivity of individual radiative components of gases, which are described in numerous references, such as [26,30,31]. For the purpose of this model, it was decided to use the mathematical model created and thoroughly described by E. Kostowski in [32,33], which was successfully applied in [29], where its validity for the gas components CO2 and H2O was confirmed.
ε i = 1 e k i . p p i . l e f n
k i = a + b . t f g 1000
εi—emissivity of the respective component, (—);
ki—correction coefficient for the effect of temperature on emissivity, (—);
p p i . l e f —the product indirectly defines the volume of the radiant component, (Pa.m);
n—exponent obtained from regression analysis, (—);
a, b—coefficients from regression analysis, (—);
tfg—flue gas temperature, (°C).
The values of coefficients a, b and the exponent n for the respective components of CO2 and H2O flue gases are determined depending on the product p p i . l e f from tables in [32,33]. The radiation of gases is not directly proportional to the fourth power of temperature. For CO2, the exponent is 3,5 ~ T3,5, and for H2O, the is exponent 3 ~ T3. For calculations, the power is chosen according to the Stefan–Boltzman law—T4 [23,26].
The necessary exponent correction is then included in the emissivity. The correction factor β is determined as follows [23,32,33]:
If p p H 2 O . l e f > 1   k P a . m (most cases), then:
β = 1 + 0.6225 0.1346 . log p p H 2 O . l e f . p p H 2 O 100 0.86
If p p H 2 O . l e f 1   k P a . m , then:
β = 1 + 0.6225 . p p H 2 O 100 0.86      
Correction of the spectral radiation of CO2 and H2O (reduction) in the total radiation of the gas mixture due to the overlap of the spectral radiation of CO2 and H2O at the same wavelengths [23,26]:
Δ ε = ε C O 2 . ε H 2 O
For the determination of heat transfer by radiation, it is very important to determine the emissivity of flue gases, which is determined from the following relationship [23,26,30]:
ε f g = ε C O 2 + β . ε H 2 O + ε S O 2 Δ ε + ε p m
ε S O 2 —SO2 emissivity, which was determined graphically from [26], (—);
ε p m —emissivity of particulate matter (dust drifts), (—).
In gaseous fuels like natural gas, sulfur is present only in trace amounts, so the hot component SO2 can be disregarded. However, during the combustion of sulfur from the reducing agent, its content may increase. Nevertheless, its content remained below 1 vol.%, and for this reason, its value in the calculations was determined graphically using a nomogram from [26].
The emissivity of particulate matter increases the overall emissivity in flue gases, and in thermal equipment, particulate matter contributes to the increase in radiant energy from the volume of flue gases. The calculations considered dust drifts during the movement of pellets through the equipment, and even after flue gas cleaning, dust particles were still present in the flue gases. The following calculation was used to determine the emissivity of particulate matter [26,30]:
ε p m = 1 e x
x = 0.75 273.15 T p m m i ρ o , p m r p m d h
m i = m ˙ p m V ˙ f g
T p m —thermodynamic temperature of particulate matter, (K);
ρ o , p m —density of particulate matter, (kg/m3);
r p m —radius of particulate matter, (m);
d h —hydraulic diameter of the free cross-section of the device, (m);
m i —particulate matter content in 1 m3 of flue gas, (kg/m3);
m ˙ p m —mass flow of particulate matter (dust drift), (kg/h);
V ˙ f g —flue gas volume flow, (m3/h).
The area density of the heat flux by radiation from the flue gas can be calculated using the following relation [23,31]:
q r = ε f g . c o . T f g 100 4 = c . T f g 100 4
ε f g —total emissivity of the flue gases (Equation (12)), (—);
c o —absolute black body radiation coefficient, 5.67 W/(m2K4);
T f g —thermodynamic temperature of flue gases, (K).
In the case of radiation from the flue gas to the charge, the previous relation (16) needs to be extended to the following form [23,31]:
q r = c . T f g 100 4 T m 100 4 = α r . Δ t
T m —thermodynamic temperature of the charge, (K);
c —grey body radiation coefficient, (W/(m2K4));
α r —radiation heat transfer coefficient, (W/(m2K));
Δ t —the temperature difference between the flue gases and the body, (°C).
(b) Radiative heat transfer between two grey bodies
Solid bodies are defined as gray bodies, meaning they cannot completely absorb or emit energy, but instead reflect part of it. Equation (17) can also be used to calculate heat transfer by radiation within a thermal unit from the walls of the unit to the charge, with the coefficient c being calculated based on Equation (19) [23,31].
q r = c . T 2 100 4 T 1 100 4 = α r . Δ t
c = c 0 1 + 1 ε 2 1 φ 21 + 1 ε 1 1 φ 12
T 1 —thermodynamic temperature of the first body, (K);
T 2 —thermodynamic temperature of the second body, (K);
ε 1 —first body emissivity, (—);
ε 2 —second body emissivity, (—);
φ 12 —the directionality of radiation from the first body to the second body, (—);
φ 21 —the directionality of radiation from the second body to the first body, (—).
The directionality of radiation is determined based on the geometry of the interior of the furnace, utilizing the rules of enclosure and reciprocity for the individual radiating surfaces. Due to the complexity of some cases, more information can be found in [23,26,30,31].
The amount of heat involved in the radiation heat exchange between two gray (real) bodies is given by the following equation [23]:
Q ˙ r = c . T 2 100 4 T 1 100 4 φ 12 S 1 = α r Δ t φ 12 S 1
Convection heat transfer
Heat transfer by convection is a part of the heat exchange not only inside the equipment but also on its surface. The convective heat exchange component depends on the nature of the flow, which also determines the choice of the appropriate mathematical relationships.
In high-temperature furnaces, the heat transferred to the material by convection typically accounts for 10–30% of the total heat transfer. This is sometimes adjusted by applying a correction factor of 1.1–1.3 to the radiant component. A more precise calculation can be achieved using Newton’s relation for surface heat flux density [23,34]:
q c = α c . Δ t
α c —convection heat transfer coefficient, (W/(m2K));
Δ t —the temperature difference of the flue gases and the body, (°C).
To determine α c , it is possible to use the approximate formula recommended by M.A. Micheev for convection heat transfer in furnaces using the dimensionless Nusselt number [23,26]:
N u = α c . L λ t
L —characteristic dimension, (m),
λ t —coefficient of thermal conductivity of the fluid, (W/(mK)).
For flow in furnaces, simplified empirical relationships can be used:
- For laminar mode:
N u = 0.57 . R e 0.5
- For turbulent mode:
N u = 0.032 . R e 0.8
For heat exchange by convection within the charge layer, mathematical models of the volumetric heat transfer coefficient α V were considered. The relationship between Newton’s coefficient α c and α V can be determined as follows:
α c = α V a s ;       ( W / ( m 2 K ) )
a s —specific surface area of the particle layer, (m2/m3).
a s = 6 1 f d ;       ( m 2 / m 3 )
f —layer porosity in the range <0;1>;
d —diameter of particles/lump, (m).
Volume coefficient of heat transfer according to Kitajev [35]:
V = A T 0.3 w o 0.7 d 0.9 M ;       ( W / ( m 3 K ) )
A—coefficient, characteristic of the material, —);
T —average gas temperature, (K);
w 0 —gas velocity, referred to 0 °C and the shaft cross-sectional area, (m/s).
d —lump diameter, (m);
M —coefficient, dependent only on bed porosity, (—).
M = 10 1.68 f 3.56 f 2 ;       ( )
Alternatively, Newton’s coefficient α c can be determined direct using a model for the dimensionless criterion Nu, as proposed by Viktorin [36]:
N u = 0.33 R e 0.77 + 0.42 ;       ( )
In the created model, Equations (27)–(29) were used in the calculation of heat transfer by convection in the drying chamber.
An important component of the heat balance is the amount of heat dissipated from the surfaces of individual devices. This heat loss is defined as free convection, but since the surface of the equipment is at a higher temperature, the radiative component from the surface to the surroundings must also be considered. For the model, overall heat transfer coefficients from the surfaces of the equipment were chosen according to Heiligenstaedt, as mentioned in [37]:
- For masonry walls:
c + r = 7.1 + 0.057 t a ;       ( W / ( m 2 K ) )
t a —external wall surface temperature, (°C).
- For metal walls:
c + r = 6.3 + 0.039 t a ;       ( W / ( m 2 K ) )
Many sources in the literature present a range of additional models for calculating heat transfer coefficients, but most are more focused on the convective component, such as those by Nusselt [37], Brunklaus [37], and others.
  • Calculation algorithm
    • Determination of the composition of leach residue;
    • Determination of the material and thermal balance of processes in the rotary kiln;
    • Determination of the energy requirements of processes in the rotary kiln: calculation of natural gas consumption and the corresponding reducing agent, as well as the residence time of material in the kiln—Rotary Kiln;
    • Determination of the material and thermal balance of processes in the cooler: calculation of the amount and temperature of cooling air and the outlet temperature of pellets—Pellet Cooler;
    • Resolution of the post-combustion process (combustion for oxide pellets) of flue gases exiting the rotary kiln—Afterburner/Combustion chamber;
    • Cooling of produced flue gases before entering the 3rd zone (preheating of raw pellets) on the Lepol grate—Flue gas Cooler I + Cogeneration (if necessary);
    • Cooling of flue gases before entering the 2nd zone (pre-drying and preheating of raw pellets) on the Lepol grate—Flue gas Cooler II;
    • Cooling of flue gases before entering the 1st zone (drying of raw pellets) on the Lepol grate—Flue gas Cooler I;
    • Calculation of flue gas temperatures before entering the filter and in the chimney.
The development of the mathematical model was based on information obtained from studies [14,16,17,18,19] that dealt with the processing of slag from Sereď, or the production of pellets from metalliferous waste:
  • The consumption of reducing agents corresponding to the direct reduction of iron oxides with carbon (C) with a coefficient of 1.5.
  • The degree of reduction of iron oxides for Sereď slag (see Figure 1). Other conditions for the mathematical model:
  • The flue gas temperature in the rotary kiln at the side where natural gas and air—necessary for burning excess carbon from the reducing agent and a portion of the released CO from the reduction, as well as volatiles from the reducing agent—are introduced should not drop below 800 °C.
  • Based on the distribution of material temperatures along the length of the kiln, the reduction of iron oxides, the burning of CO and volatiles, and the burning of excess carbon from the reducing agent were determined.
Since the determining process for thermal calculations is the heating and reduction of iron oxides in the rotary kiln, parts of the mathematical model that addressed processes in the rotary kiln and subsequently in the pellet cooler were treated zonally to obtain temperature profiles of the flue gases and the charge along the length of these units.

3. Results

The modelling results are based on a calculation algorithm that utilizes modules for material and energy balance calculations in the individual units. The obtained results are presented in tables and graphs.

3.1. Material Balance of Pellet Composition

Based on the composition of the raw pellets and the amount of reducing agent, the mathematical model calculations included a recalculation of the pellet composition after thermal processing and cooling. The output pellet composition is the final composition and is provided for each calculation variant in Table 4.
The amount of reducing agent in the raw pellets corresponds to the amount of carbon in Table 4. The results after reduction show that in no case was the carbon completely consumed for reduction when using the reducing agent. The ballast substances from the reducing agent are reflected in the ash content, and as can be observed, the value is higher when using lignite compared to coke. The input ratio of ash in lignite to coke is, on average, 2.71, and at the output, it is 2.99. The average moisture content in the raw pellets was approximately 13.372 wt.%, with the oxide pellets having an increased moisture content of 8 wt.%.

3.2. Outputs of Calculations for Individual Equipment

The results from the calculations using the developed mathematical model, which formed the basis for the overall material and energy balance, are presented in Table 5. The individual units, which are sequentially linked in the raw pellet processing process, are listed progressively.
The results of the calculations for various alternatives indicate the overall energy consumption of the entire processing operation. The data show that achieving a 92% reduction of oxides in pellets using coke requires 3.87 GJ/t of pellets more energy compared to a 61% reduction. For lignite, achieving a 69% reduction of oxides requires 2.821 GJ/t of pellets more energy, which on average represents a 40% increase in energy consumption compared to the energy required for a 92% reduction of oxides using both coke and lignite. Therefore, it is essential to assess the practical feasibility of these alternatives.

3.3. Pellet Cooling

For all alternatives, a constant amount of air is considered, which has been determined to prevent or limit the reoxidation of the produced pellets. Figure 5 shows the temperatures of the pellets and the cooling air. The cooler is designed as a counterflow heat exchanger, where the pellet temperature decreases from approximately 1100 °C to 410 °C, and the air is heated to about 380 °C.
This type is the most efficient for maximizing the cooling of the pellets using a constant amount of air while simultaneously achieving a high temperature for the preheated air.

3.4. Basic Calculation Parameters of Processes in a Rotary Kiln

The necessary input calculations for each alternative were conducted using a developed mathematical model, which was used to establish the overall material and energy balances in the rotary kiln. The results of these calculations are presented through graphs and summarized in tables that illustrate the performance of each alternative. This chapter provides a detailed description of these alternatives, their calculations, and the analysis of results for the rotary kiln.

3.4.1. Reducing Agent: Coke—50% Reduction

To achieve the minimum amount of total iron in the pellet, a reduction with coke at the 50% reduction boundary of oxides was proposed. Figure 6a shows the temperature profiles of the exhaust gases and pellets along the length of the rotary kiln. The calculations revealed that the residence time of the pellets in the rotary kiln is 103 min.
The raw pellets, after entering the rotary kiln, are heated to the required reduction temperature of at least 950 °C approximately 17 m into the kiln, corresponding to a residence time of 43 min. During the movement of the raw pellets, reduction and the combustion of carbon within them occurs. This affects the temperature of the raw pellets and the exhaust gases along the rotary kiln.
The expected course of excess carbon combustion, CO combustion from reduction, and volatile matter combustion can be observed in Figure 6b.
Analyses conducted for the reduction of iron ore pellets indicate that, in addition to temperature, the residence time of the pellets within these temperature zones is crucial. Reduction begins at temperatures of approximately 950 °C. Figure 6a,b show that at 50% reduction, the material remains in the kiln at temperatures above 950 °C for approximately 60 min. Experiments suggest that the time required for approximately 50% reduction of iron oxides is around 50 min.
From the production requirements, it follows that for the alternative of 50% reduction of iron oxides with coke, a feedstock with the following mass flow or ratio of individual raw materials per ton of produced pellets is necessary, as presented in Table 6.

3.4.2. Reducing Agent: Coke—61% Reduction

The minimum threshold of 50% reduction of oxides in raw pellets was determined based on experimental measurements and calculations related to the reactivity of coke, considering the degree of reduction and residence time in the kiln. Based on these calculations from sources [18,19], a second threshold was also established for the variant with a 61% reduction level using coke.
The results of the temperature profile calculations along the length of the kiln are shown in Figure 7a. Similar to the previous variant, in this case, the combustion of carbon in the pellets begins once the pellets reach a sufficient temperature, and complete combustion is calculated to occur at the end of the rotary kiln (Figure 7b).
Figure 7a indicates that in the final meters of the rotary kiln, the raw pellets are not heated by the flue gases but by the reactions occurring within them and the combustion of coke. However, the flue gases are crucial for achieving the required temperature of the raw pellets to initiate and sustain the reduction of Fe oxides.
Compared to the 50% reduction with coke, there is no significant difference in the residence time of the raw pellets in the rotary kiln or their temperature. The residence time is approximately 60 min, and experiments indicate that the time required for about a 60% reduction of iron oxides is around 55–60 min.

3.4.3. Reducing Agent: Coke—92% Reduction

To determine the degree of Fe oxide reduction using coke as a reducing agent, calculations were conducted similarly to the previous variant. This alternative was chosen to maximize the reduction of iron oxides from the original material. The results of the temperature profiles of the flue gases and pellet temperatures are shown in Figure 8a.
From Figure 8a, it can be observed that the residence time of the raw pellets at temperatures above 950 °C is approximately 60 min. Experiments indicate that the time required to achieve an Fe oxide reduction of over 90% is 180 min. This parameter can be achieved by adjusting the feed rate in the rotary kiln.
Table 6 presents the input parameters of the raw pellets for each variant using coke as a reducing agent. The calculations were performed for the material and energy balance, considering mass and volumetric flows and converted per ton of material according to the relevant sections.
The energy consumption increases in a slightly parabolic manner as the degree of Fe oxide reduction increases. Compared to the maximum reduction level of 92%, the energy consumption increases by 51.1% over the chosen minimum reduction level and by 39.4% over a 61% reduction level. These results demonstrate the significant rise in required energy in the raw pellet processing. Energy consumption was selected as one of the criteria for evaluating the practical feasibility of each alternative.

3.4.4. Reducing Agent: Lignite—69% Reduction

As an alternative reducing agent, lignite was defined, which differs significantly from coke in terms of volatile combustible content, ash content, and moisture. The degree of Fe oxide reduction was determined based on experimental measurements [18,19] and calculations that were derived from the reactivity of lignite, the degree of Fe oxide reduction, and the residence time in the kiln. In this case, the minimum degree of Fe oxide reduction was set at 69%.
As the analysis shows, the rotary kiln can reach the temperatures required for reduction; however, the raw pellets only achieve these temperatures in the middle of the kiln, significantly shortening the time the raw pellets spend at temperatures above the minimum reduction temperature of approximately 950 °C.
From Figure 9a, it is evident that the residence time of the raw pellets at temperatures above 950 °C is around 40 min. Experiments indicate that the time required to achieve approximately 70% reduction of Fe oxides is about 70–80 min. This implies that to achieve a higher degree of reduction, either a longer kiln or a slower feed rate would be necessary.

3.4.5. Reducing Agent: Lignite—92% Reduction

The chosen reduction level of 92% Fe oxides represents the maximum possible reduction of iron oxides using lignite. The results of the temperature profile calculations for the flue gases and pellets along the length of the kiln are shown in Figure 10a.
Again, Figure 10a shows the short residence time of the raw pellets at temperatures above 950 °C, which is approximately 40 min. Experiments referenced in [18,19] indicate that the time required for over 90% reduction of iron oxides is 180 min.
Table 7 presents the calculated energy consumption values for the variant using lignite as the reducing agent. The results show an increase of 2.82 GJ/t of pellets, corresponding to a 28.65% increase in energy consumption.

3.4.6. Oxide Pellets: Without Reduction of Fe Oxides

Another option for utilizing the leach residue is the production of pellets without reducing agents. In this case, the reduction of iron oxides does not occur; rather, the process focuses on creating conditions or improving the properties of the pellets (abrasion resistance, absorbency, strength, etc.) for their further use.
The temperature profile of the raw pellet heating process, shown in Figure 11, illustrates a steady increase in pellet temperature depending on the heating regime in the rotary kiln. Heat transfer occurs through convection, radiation, and, most significantly, conduction between the pellets and the inner wall of the rotary kiln. The temperature of the pellets may be influenced by internal phase transformations within the pellet composition, though this change is negligible.
Natural gas is used for each stage of the process. For the production of oxide pellets, the residence time of the raw pellets in the rotary kiln (due to the absence of reduction) is less important; the primary concern is reaching the required pelletizing temperature.
As shown in Table 8, energy consumption is the lowest when compared to the energy consumption of other alternatives that use reducing agents. This alternative appears to be the most suitable from an energy efficiency standpoint. However, attention must be paid to the quality (% Fe in the pellet) and the cost of energy commodities. Currently, prices are changing dynamically depending on demand but especially due to the global social situation and extraordinary events around the world.

4. Discussion

Thermodynamic calculations indicate that it is possible to produce pellets with a reduction degree of iron oxides from the Sereď leach residue in the metallurgical facilities of Nižná Slaná, achieving 50–65% reduction. This corresponds to a residence time of the feedstock in the temperature zone of 950–1100° C for approximately 50–65 min (Figure 6a). A higher reduction degree is not feasible or would be inefficient due to technological constraints.
The calculations for the material’s residence time in the rotary kiln were based on a kiln length of 40 m, a kiln inclination of 3%, and a rotation speed of 0.5 revolutions per minute. The residence time in the kiln is approximately 103 min. At higher rotation speeds, the residence time in the kiln would decrease.
In the production of reduced pellets using coke as a reducing agent, it is possible to achieve a total Fe content of 60% in the produced pellets at a 50% reduction of iron oxides (Figure 1 and Figure 6b). In the Nižná Slaná technology, it is also possible to achieve a reduction of iron oxides to 62%, which would increase the total Fe content in the pellets to approximately 62% (Figure 1 and Figure 7a). Theoretically, the maximum increase in total Fe content in the pellets would be around 65% at a 92% reduction of iron oxides in the feedstock (Figure 1 and Figure 7b), but this alternative is unrealistic under the conditions of the facilities in Nižná Slaná and is likely economically inefficient.
In the production of reduced pellets using lignite as a reducing agent, it is possible to achieve a total Fe content of 60% in the produced pellets with a 69% reduction of iron oxides (Figure 1 and Figure 9a). However, this is beyond the operational limits of the Nižná Slaná technology in terms of the required residence time of the material in the kiln at the necessary temperatures for the reduction of iron oxides. When using lignite as a reducing agent, raw pellets can be preheated on the Lepol grate to a maximum temperature of 400 °C. At higher temperatures, volatile substances start to release, which cannot be economically burned before entering the chimney (due to the large volume of exhaust gases, 200,000–300,000 m3/h, at a relatively low temperature of around 200 °C).
A disadvantage of lignite as a reducing agent is its composition; in addition to high moisture content, it also has a high ash content, which remains in the pellets after reduction and reduces the total Fe content. Therefore, even with a 92% reduction of iron oxides, the total Fe content in the pellet only increases to about 62% (Table 4). The moisture content of the lignite also plays a crucial role; lignite with a moisture content higher than 23.6% should not be added to the feedstock. Higher moisture content degrades the quality of the feedstock and increases energy consumption for drying.
In the case of oxide pellet production, the process from a thermal–technical perspective involves only the drying and heating of the feedstock, with structural changes occurring at higher temperatures from a technological standpoint. The energy demand for this process is determined by the consumption of natural gas (Table 8).
Key data from the energy evaluation for various production scenarios of the feedstock from the Sereď leach residue (Fe concentrate SE552) are presented in Table 9.
Cooling the produced pellets with air in a cooler should be carefully considered, as it may lead to partial reoxidation of the reduced pellets. An alternative cooling method, such as indirect water cooling, should be explored. However, this would increase the consumption of natural gas in the rotary kiln by approximately 650 m3/h or increase the energy demand of the process for 50% reduction with coke to 5.32 GJ/t (an increase of 0.49 GJ/t) and would also require the establishment of a water management system.
A potential drawback of the rotary kiln process could be the introduction of combustion air for burning residual carbon during direct reduction and the combustion of a portion of CO from the reduction, as well as H2 and hydrocarbons from the volatiles in the reducing agent. The most advantageous approach is to ensure the supply of air into the kiln’s working chamber from two to three locations (kiln entry—gas supply side; at one-quarter of the kiln; at half of the kiln). This would help to even out the combustion of carbon, CO, and volatiles in the kiln and ensure higher flue gas temperatures at the flue gas entry side, which could also lead to a partial reduction in natural gas consumption.
In the case of utilizing waste heat, it is possible to use cogeneration units without combustion, only with heat exchangers, due to the potential contamination of the exhaust gases from the afterburner. Two solutions were proposed for this scenario:
The first solution involves placing the cogeneration unit behind the mixing chamber located after the afterburner. For this case, an ORC-type cogeneration unit would be suitable. These units operate at lower input temperatures due to the properties of the heat transfer medium.
The second solution considers placing a cogeneration unit with indirect heating directly behind the afterburner. Such a cogeneration unit operates as an open Brayton cycle, using air as the working medium, which, after expansion, can be further utilized in the technological process, for example, as preheated air for the afterburner or the mixing chamber.
If we were to compare energy consumption in similar operations, for example, the study [38] reports the minimum energy consumption for selected processes in steel production. The authors indicate the necessary minimum energy, and for the reduction of iron oxides, they calculated an energy consumption of 8.6 GJ/t. In another work [39], the authors describe the energy consumption during the pelletizing process. The overall energy consumption for pellet production, compared to study [38], is particularly interesting, as it is only 793.4 MJ/t of pellets produced. In another study [40], the authors compared energy consumption in the production of pellets from hematite and magnetite, where the energy consumption ranged between 1.2 and 1.7 GJ/t of pellets produced.
These studies indicate that it is possible to reduce energy consumption and thereby operational costs by introducing more modern technologies or by modernizing existing ones. However, it is very difficult to compare each technology fully due to the specificity of the equipment and input materials, which makes it necessary to analyze the possibilities for modifications or the introduction of more environmentally friendly and energy-efficient technologies. The answer to these challenges lies in the energy and material balance of the entire process, which was also the authors’ aim in this article.
For comparing energy consumption in similar operations, it is essential to consider the processing method and the equipment used. Comparing alternatives with a reduction degree of 50–69%, energy consumption ranges from 4.835 to 7.023 GJ/t of pellets produced. In the production of oxide pellets, where no reducing agent was used, the total energy consumption is only 2.167 GJ/t of pellets produced. This suggests that the described technology in Nižná Slaná has the potential for producing pellets from metal-bearing waste.

5. Conclusions

The authors conducted a material and energy evaluation of the process for processing raw pellets made from waste material that has not yet been utilized. In line with environmental policy goals for the years 2030 and 2050, European countries are interested in addressing such environmental issues.
One advantage of the waste from the Sereď leach residue dump is its high Fe content, which can be utilized in the metallurgical industry either as sinter or pellets. The study of the energy assessment for processing raw pellets from the Fe concentrate SE52 leach residue demonstrated the feasibility of processing these materials in the decommissioned Nižná Slaná metallurgical plant. The calculations were based on the actual parameters of the existing equipment, which could be repurposed for waste disposal and utilization. The advantage lies in the existence of equipment that could be brought back into operation with additional investments in renovation and environmental upgrades.
The calculation results showed that from an energy and process perspective, it is feasible to produce pellets with a reduction degree of Fe oxides between 50 and 65%. The rotary kiln parameters proved to be suitable for processing raw pellets from the Fe concentrate SE52 leach residue. Coke and lignite were used as reducing agents for direct reduction in the raw pellets, and their suitability was confirmed by calculations. The high carbon content in these reducing agents—69.95% in coke and 41.36% in lignite—contributed to this suitability. However, the high moisture content in lignite was identified as a challenge, as it adversely affects the processing of raw pellets and increases the energy demands of the entire process. A higher degree of Fe oxide reduction can be achieved at the cost of increased energy consumption, with a 92% reduction requiring 40% more energy compared to a 61% reduction with coke and a 69% reduction with lignite.
Experiments with leach residue samples showed that the degree of Fe oxide reduction could be further increased by changing the reducing agent to hydrogen (Figure 1). The research on using different reducing agents is ongoing, with oil showing promise due to its high carbon content of 76.95%. However, a drawback of using oil as a reducing agent is its rapid evaporation in the drying chamber, which would require redesigning the exhaust system for afterburning volatiles from the oil.
This study fulfilled its purpose, and the recommendations based on the material and energy balance indicate the feasibility in reviving the decommissioned technology. The processed pellets are intended for use as part of the charge in the production of pig iron at a nearby metallurgical plant.

Author Contributions

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

Funding

This article was supported by the Cultural and Educational Grant Agency of the Ministry of Education, Research, Development, and Youth of the Slovak Republic, through the project KEGA 024TUKE-4/2024.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pavolová, H.; Čulková, K.; Šimková, Z.; Seňová, A.; Kudelas, D. Contribution of Mining Industry in Chosen EU Countries to the Sustainability Issues. Sustainability 2022, 14, 4177. [Google Scholar] [CrossRef]
  2. European Environment Agency. Resource Efficiency and Waste. Available online: https://www.eea.europa.eu/sk/themes/waste/intro (accessed on 4 June 2024).
  3. Environment. (n.d.). Environment Action Programme to 2030. Available online: https://environment.ec.europa.eu/strategy/environment-action-programme-2030_en?prefLang=sk&etrans=sk (accessed on 6 June 2024).
  4. Majerčák, Š. Pelletization of Fine-Grained Materials; Alfa: Bratislava, Slovakia, 1977. [Google Scholar]
  5. Mihok, Ľ.; Frohlichová, M.; Baricová, D.; Pribulová, A. Fundamentals of Metallurgical Technologies; TU of Košice: Košice, Slovakia, 2004. [Google Scholar]
  6. Laubertova, M.; Ruzickova, S.; Trpcevska, J.; Briancin, J. Experimental Study of a Representative Sample to Determine the Chemical Composition of Cast Iron. Materials 2024, 17, 1255. [Google Scholar] [CrossRef]
  7. Fan, X.; Li, J.; Chen, X.; Wang, Y.; Gan, M. Temperature Field Simulation Model for Rotary Kiln of Iron Ore Oxidized Pellet. J. Iron Steel Res. Int. 2013, 20, 16–19. [Google Scholar] [CrossRef]
  8. Larsson, I.A.S. The Aerodynamics of an Iron Ore Pelletizing Rotary Kiln. Fluids 2022, 7, 160. [Google Scholar] [CrossRef]
  9. Gorog, J.P.; Adams, T.N.; Brimacombe, J.K. Heat transfer from flames in a rotary kiln. Metall. Trans. B-Process Metall. 1983, 14, 411–424. [Google Scholar] [CrossRef]
  10. Davis, R.A. Mathematical model of magnetite oxidation in a rotary kiln furnace. Can. J. Chem. Eng. 1996, 74, 1004–1009. [Google Scholar] [CrossRef]
  11. Sadrnezhaad, S.K.; Ferdowsi, A.; Payab, H. Mathematical model for a straight grate iron ore pellet induration process of industrial scale. Comput. Mater. Sci. 2008, 44, 296–302. [Google Scholar] [CrossRef]
  12. Constantin, N.; Socalici, A.V.; Buzduga, R.; Mihaiu, A.C.; Dobrescu, C.; Stanasila, O.N.; Plopeanu, E.-F.; Vlad, E.M. Experimental research on a semi-industrial pilot scale for obtaining carbofer pellets in rotary tubular furnace. Univ. Politeh. Buchar. Sci. Bull. Ser. B-Chem. Mater. Sci. 2021, 83, 155–164. [Google Scholar]
  13. Wang, G.; Wang, J.S.; Xue, Q.G. Efficient utilization of waste plastics as raw material for metallic iron and syngas production by combining heat treatment pulverization and direct reduction. Process Saf. Environ. Prot. 2020, 137, 49–57. [Google Scholar] [CrossRef]
  14. Imriš, I. Technical Report: Technological Research on the Processing of Fe-Concentrate from Sered’ with the Possibility of Its Utilization by 2010; TU of Košice: Košice, Slovakia, 1998. [Google Scholar]
  15. Fan, W.; Peng, Z.; Tian, R.; Luo, G.; Yi, L.; Rao, M. Toward Metallized Pellets for Steelmaking by Hydrogen Cooling Reduction: Effect of Gas Flow Rate. Materials 2024, 17, 3896. [Google Scholar] [CrossRef]
  16. Kudělka, Z.; Drabina, J.; Vítek, V.; Schmidt, H.; Benoni, V. Processing of pellets from Albanian leachate residue in a blast furnace. Part I. Hut. Listy/Metall. J. 1985, 40, 686–694. [Google Scholar]
  17. Kudělka, Z.; Drabina, J.; Vítek, V.; Schmidt, H.; Benoni, V. Processing of pellets from Albanian leachate residue in a blast furnace. Part II. Hut. Listy/Metall. J. 1985, 40, 762–767. [Google Scholar]
  18. Růžičková, Z.; Srb, J.; Hubner, A.; Žádný, I. Possibilities of metallurgical use of modified Albanian leachate residue. Hut. Listy/Metall. J. 1986, 41, 1–18. [Google Scholar]
  19. Míček, M.; Repášová, M.; Varga, A. Tubular rotary kiln for pellet pre-reduction. Hut. Listy/Metall. J. 1987, 42, 305–309. [Google Scholar]
  20. Tomášek, K. Defunct Metallurgical Technologies of Non-Ferrous Metals in Slovakia—Production of Nickel and Cobalt in Slovakia. Hut. Listy/Metall. J. 2017, LXX, 25–27. [Google Scholar]
  21. SPP-Distribucia: Natural Gas Composition and Emission Factor. Available online: https://www.spp-distribucia.sk/dodavatelia/informacie/zlozenie-zemneho-plynu-a-emisny-faktor/ (accessed on 4 March 2017).
  22. Durcansky, P.; Papucik, S.; Jandacka, J.; Holubcik, M.; Nosek, R. Design of Heat Exchanger for Ericsson-Brayton Piston Engine. Sci. World J. 2014, 2014, 138254. [Google Scholar] [CrossRef]
  23. Varga, A.; Jablonský, G.; Lukáč, L.; Kizek, J. Thermal Technology for Metallurgists, 1st ed.; Technical University: Košice, Slovakia, 2013. [Google Scholar]
  24. Prokopchuk, M.; Kizek, J.; Varga, A. Effect of flue gas recirculation and oxygen addition on combustion parameters in the Brayton cycle. Adv. Therm. Process. Energy Transform. 2023, 6, 32–39. [Google Scholar] [CrossRef]
  25. Komorova, L.; Imris, I. Thermodynamics in Metallurgy, 2nd ed.; Alfa: Bratislava, Slovakia, 1992. [Google Scholar]
  26. Rédr, M.; Příhoda, M. Fundamentals of Thermal Engineering, 1st ed.; SNTL: Prague, Czech Republic, 1991. [Google Scholar]
  27. Sazima, M.; Kmoníček, V.; Schneller, J.; Ambroz, J.; Bayer, Z.; Cerny, V.; Hlavacka, V.; Pikman, M.; Streda, I.; Sifner, O. Heat, 1st ed.; SNTL: Prague, Czech Republic, 1989. [Google Scholar]
  28. Cengel, Y.; Ghajar, A. Heat and Mass Transfer: Fundamentals and Applications, 6th ed.; McGraw Hill: New York, NY, USA, 2019. [Google Scholar]
  29. Rimar, M.; Kizek, J.; Varga, A.; Fedak, M.; Jablonský, G. The influence of hydrogen concentration in natural gas on heat flows in a thermal aggregate. MM Sci. J. 2022, 6162–6168. [Google Scholar] [CrossRef]
  30. Hottel, H.C.; Sarofim, A.F. Radiative Transfer, 1st ed.; McGraw-Hill Book Co.: New York, NY, USA, 1967. [Google Scholar]
  31. Ferstl, K.; Masaryk, M. A Collection of Examples from Heat Transfer—Heat Transfer by Radiation and Evaporation of Water; SPEKTRUM STU: Bratislava, Slovakia, 2018. [Google Scholar]
  32. Kostowski, E. Approximation of Charts Describing CO2 and H2O Emission Using Analytical Functions (Aproksymacja Wykresów Określających Emisyjność CO2 i H2O, za Pomocą Funkcji Analitycznych); s. Energetyka, 92; Zeszyty Naukowe Politechniki Śląskiej: Gliwice, Poland, 1987. [Google Scholar]
  33. Pronobis, M. Accuracy of Calculations of Radiative Heat Transfer from Flue Gas to the Surface of Boiler Tubes of Convection Bundles (Dokladnosc Obliczen Radiacyjnej Wymiany Ciepla od Spalin do Powierzchni rur Kotlowych Pęczkow Konwekcyjnych); Energetyka, 114; Zeszyty Naukowe Politechniki Slaskiej: Gliwice, Poland, 1993. [Google Scholar]
  34. Ferstl, K.; Masaryk, M. A Collection of Examples from Heat Transfer—Heat Transfer by Convection; STU: Bratislava, Slovakia, 2016. [Google Scholar]
  35. Kitajev, B.I.; Yaroshenko, Y.G.; Suchkov, V.D. Heat Exchange in Shaft Furnaces, 1st English ed.; Pergamon Press, Ltd.: Oxford, UK; New York, NY, USA, 1967. [Google Scholar]
  36. Viktorin, Z. Heat Transfer in the Flow through a Stationary Bed of Grain Material. Zdravotní technika a vzduchotechnika/Med. Technol. Air Cond. 1968, 11, 225–234. [Google Scholar]
  37. Redr, M.; Gottwald, M.; Riman, A.; Rejc, R. Thermal Calculations and Optimization of Refractory Linings in Industrial Furnaces; SNTL: Prague, Czech Republic, 1975. [Google Scholar]
  38. Fruehan, R.J.; Fortini, O.; Paxton, H.W.; Brindle, R. Theoretical Minimum Energies to Produce Steel for Selected Conditions; Carnegie Mellon University: Pittsburgh, PA, USA; U.S. Department of Energy Office of Industrial Technologies: Washington, DC, USA, 2000.
  39. Lv, W.; Sun, Z.; Su, Z. Life cycle energy consumption and greenhouse gas emissions of iron pelletizing process in China, a case study. J. Clean. Prod. 2019, 233, 1314–1324. [Google Scholar] [CrossRef]
  40. Hooey, L.; Riesbeck, J.; Wikström, J.-O.; Björkman, B. Role of Ferrous Raw Materials in the Energy Efficiency of Integrated Steelmaking. ISIJ Int. 2014, 54, 596–604. [Google Scholar] [CrossRef]
Figure 1. Dependence of Fe reduction efficiency on the time and type of reducing agent used at a temperature of 1100 °C (adapted from [3]).
Figure 1. Dependence of Fe reduction efficiency on the time and type of reducing agent used at a temperature of 1100 °C (adapted from [3]).
Processes 12 01938 g001
Figure 2. Technological equipment for pellet processing in Nižná Slaná, Slovakia.
Figure 2. Technological equipment for pellet processing in Nižná Slaná, Slovakia.
Processes 12 01938 g002
Figure 3. Technological scheme of the main parts of the pellet production process equipment for the alternative without cogeneration.
Figure 3. Technological scheme of the main parts of the pellet production process equipment for the alternative without cogeneration.
Processes 12 01938 g003
Figure 4. Technological scheme of the main parts of the pellet production process equipment for the alternative with cogeneration.
Figure 4. Technological scheme of the main parts of the pellet production process equipment for the alternative with cogeneration.
Processes 12 01938 g004
Figure 5. Temperature of pellets and cooling air along the length of the cooler.
Figure 5. Temperature of pellets and cooling air along the length of the cooler.
Processes 12 01938 g005
Figure 6. Alternative C_50R for reduction with coke at 50% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Figure 6. Alternative C_50R for reduction with coke at 50% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Processes 12 01938 g006
Figure 7. Alternative C_61R for reduction with coke at 61% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Figure 7. Alternative C_61R for reduction with coke at 61% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Processes 12 01938 g007
Figure 8. Alternative C_92R for reduction with coke at 92% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Figure 8. Alternative C_92R for reduction with coke at 92% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Processes 12 01938 g008
Figure 9. Alternative BC_69R for reduction with coke at 69% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Figure 9. Alternative BC_69R for reduction with coke at 69% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Processes 12 01938 g009
Figure 10. Alternative BC_92R for reduction with coke at 92% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Figure 10. Alternative BC_92R for reduction with coke at 92% reduction level: (a) temperature of the raw pellets along the length and residence time in the rotary kiln; (b) percentage of reduction and combustion of CO, H2, and volatiles; combustion of excess C; and their overlap along the length and residence time in the rotary kiln.
Processes 12 01938 g010
Figure 11. Temperature of the raw pellets along the length in the rotary kiln (oxidized pellets).
Figure 11. Temperature of the raw pellets along the length in the rotary kiln (oxidized pellets).
Processes 12 01938 g011
Table 1. Composition of leach residue of Fe concentrate SE52.
Table 1. Composition of leach residue of Fe concentrate SE52.
Composition of Fe ConcentrateAverage Values for Calculations
Min. wt.%Max. wt.%wt.%
Fe total from oxides48.0052.00
Fe2O342.0043.0045.539
FeO27.0028.0029.466
Fe metal0.250.350.3214
SiO28.0010.009.643
Al2O34.006.005.357
CaO3.004.504.018
MgO2.003.002.679
Cr*1.301.501.500
P2O50.060.180.129
SO30.080.100.096
MnO0.300.400.375
K2O0.080.100.096
Ni0.270.290.300
CoO0.0620.0620.066
Cu0.010.020.016
Zn0.030.030.032
V2O50.020.020.021
TiO20.100.120.118
Bi0.0010.010.006
Na2O0.200.210.220
H2O5.0018.0012.322
Σ93.71115.82112.322
Cr*—in the form Cr2O3.
Table 2. Composition and calorific value of the proposed reducing agents.
Table 2. Composition and calorific value of the proposed reducing agents.
ComponentCSHONAshWaterLHV
wt.%wt.%wt.%wt.%wt.%wt.%wt.%MJ/kg
Coke69.950.371.400.700.778.8118.0023.56
Lignite41.361.843.1615.930.6713.4523.6015.44
Table 3. Composition and calorific value of natural gas [21].
Table 3. Composition and calorific value of natural gas [21].
Componentvol. %Componentvol. %
CH496.858izo,n—C5H120.022
C2H61.421C6H14 + higher0.026
C3H80.444CO20.212
izo,n—C4H100.139N20.878
LHV *36.329 MJ/m3
* The value is given under normal conditions of temperature 0 °C and pressure 101,325 Pa.
Table 4. Composition of raw pellets at the inlet and pellets at the outlet of the process.
Table 4. Composition of raw pellets at the inlet and pellets at the outlet of the process.
Reducing AgentsCokeLigniteOxidized Pellets
Degree of Reduction50%61%92%69%92%0%
InputOutputInputOutputInputOutputInputOutputInputOutputInputOutput
wt.%wt.%wt.%wt.%wt.%wt.%
Fe total 60.339 61.670 65.525 60.444 62.177 55.244
Fe0.24421.3460.23831.0140.22459.0100.21336.4070.20055.9950.2730.320
FeO22.49750.13421.97739.41620.7028.37619.67930.90618.4757.94825.23929.530
Fe2O334.7790.00033.9750.00032.0030.00030.4230.00028.5610.00039.01945.652
CaO3.0404.2862.9704.3802.7984.6542.6604.2932.4974.4163.4113.924
SiO27.32910.5087.16010.7406.74411.4116.41110.5266.01910.8288.2239.621
MgO2.0272.9061.9802.9701.8653.1561.7732.9111.6652.9952.2742.661
C *8.6200.20610.3920.26014.7390.41610.4820.28713.1530.7890.0000.000
Al2O34.0545.8123.9605.9403.7306.3123.5465.8223.3295.9894.5485.321
P2O50.0960.1370.0940.1400.0880.1490.0840.1380.0790.1420.1080.126
CaOSO30.0730.1790.0720.1830.0670.1940.0640.1790.0600.1840.0820.164
K2O0.0730.1050.0720.1070.0670.1140.0640.1050.0600.1080.0820.096
ZnO (s)0.0250.0350.0240.0360.0230.0380.0210.0350.0200.0360.0280.032
Ni(NiO)0.2290.3280.2240.3350.2110.3560.2000.3290.1880.3380.2570.300
CoO0.0120.0170.0120.0180.0110.0190.0110.0170.0100.0180.0140.016
V2O5(V2O3)0.0160.0230.0160.0240.0150.0250.0140.0240.0130.0240.0180.021
TiO20.0900.1290.0880.1310.0830.1400.0780.1290.0740.1320.1010.118
MnO0.2850.4080.2780.4170.2620.4430.2490.4090.2340.4200.3190.374
Cr2O5(Cr2O3)1.1421.6381.1161.6741.0511.7780.9991.6410.9381.6881.2821.499
Bi(Bi2O3)0.0040.0060.0040.0060.0040.0070.0040.0060.0040.0060.0050.006
Na2O0.1680.2400.1640.2460.1540.2610.1470.2410.1380.2480.1880.220
Ash1.0861.5561.3091.9631.8563.1413.4095.5964.2787.6950.0000.000
N2 + O2 + H2 + S0.3990.0000.4810.0000.6830.0005.4740.0006.8690.0000.0000.000
H2O13.7130.00013.3960.00012.6190.00013.9950.00013.1380.00014.5300.000
* C recalculated from the reducing agents.
Table 5. Outputs from the mathematical model for the different variants of reduction and reducing agent in the individual process equipment.
Table 5. Outputs from the mathematical model for the different variants of reduction and reducing agent in the individual process equipment.
Alternative C_50RC_61RC_92RBC_69RBC_92ROX
Fuel for the rotary kiln
Natural gas (NG) (m3/kgfuels) 1.3511.3511.3511.3511.3511.351
LHV (kJ/kgfuels) 49,09649,09649,09649,09649,09649,096
Excess air coefficient for combustion NG (-) 1.11.11.11.11.11.1
Mass flow of wet charge (kg/h) 69,69472,91682,246.679,81187,45256,875
Cooler
Mass flow of charge (kg/h)input47,63947,63947,63947,63947,63947,639
output47,63947,63947,63947,63947,63947,639
Temperature of charge (°C)input110011001100110011001100
output410410410410410410
Volumetric air flow (m3/h) 35,00035,00035,00035,00035,00035,000
Air temperature (°C)input202020202020
output380380380380380380
Heat supplied by the charge (kJ/h) 28,83728,36426,99128,28527,50431,004
Heat fluxes per charge (kJ/h) 27,55927,55927,55933,84833,84833,848
Useful heat—air (kJ/h) 16,77816,77816,77816,77816,77816,778
Rotary kiln
Mass flow of charge (kg/h)input60,13763,14871,86868,64275,96248,611
output47,63947,63947,63947,63947,63947,639
Temperature of charge (°C)input650650650400400500
output110011001100110011001100
Flue gas temperature (°C)input8508448378378832057
output97310061051904955643
Volumetric air flow—total (m3/h) 34,16743,48173,22736,49449,65014,831
Preheated air temperature (°C) 380380380380380380
Mass flow of fuel (kg/h) 51874014805559621036
Fuel volume flow (m3/h) 7001000200075013001400
Flue gas volume flow (fuel) (m3/h) 812511,60723,214870515,08916,250
Flue gas volume flow—total (m3/h) 42,13253,82490,28554,36974,80316,250
Flue gas composition (vol.%)CO10.72910.60510.9768.9999.7600.000
CO217.14717.13915.93820.65119.7628.791
H2O3.9753.7305.0492.7693.48917.297
SO20.0530.0520.0500.0000.0000.000
N264.20563.95964.21153.23952.65172.170
O21.3351.3601.3681.1931.1981.742
Degree of reduction 5061926992.30
Degree of free C combustion 0.950.950.950.950.90
Degree of C combustion to CO2 111110
Degree of CO Combustion from reduction 0.400.400.350.530.490
Mass flow of dust drifts (kg/h) 972972972972972972
Mass flow of dust drifts—total (kg/h) 972972972972972972
Flue gas velocity in the rotary kiln (m/s) 6.4948.27513.9225.2787.4151.906
Residence time of flue gases in the rotary kiln (s) 3.0802.4171.4370.7580.53910.493
Heat supplied by the charge (kJ/h) 147,291182,418292,217204,316270,67851,245
Heat fluxes per charge (kJ/h) 22,51520,53220,23639,45343,89427,542
Useful heat of charge (kJ/h) 78,14392,380133,519116,831147,66728,159
Heat fluxes required per charge (kJ/h) 22,85520,73622,85537,87842,80328,159
Specific fuel consumption (m3/t) 14.721.042.015.727.329.4
Cyclone
Fuel volume flow (m3/h)input42,13253,82490,28554,36974,80316,250
Flue gas temperature (°C)input97310061051904955643
Flue gas composition (vol.%)H22.563.162.419.009.7600.000
CO10.72910.60510.9768.9999.7600.000
CO217.14717.13915.93820.65119.7628.791
H2O3.9753.7305.0492.7693.48917.297
SO20.0530.0520.0500.0000.0000.000
N264.20563.95964.21153.23952.65172.170
O21.3351.3601.3681.1931.1981.742
Mass flow of dust drifts (kg/h)output972972972972972972
Mass flow of dust drifts—total (kg/h)output194194194194194194
Fuel composition—combustion chamber
Natural gas(NG) (kg/kgfuels) 111111
Natural gas(NG) (m3/kgfuels) 1.3511.3511.3511.3511.3511.351
LHV (kJ/kgfuels) 49,09649,09649,09649,09649,09649,096
Afterburner chamber (DSK)
Mass flow of fuel (kg/h) 14800373371110
Fuel volume flow (m3/h) 2000050501500
Volumetric flow rate of CO + H2 afterburning (m3/h) 55971698217412,04117,1300
Volumetric combustion air flow (m3/h) 16,99619,39831,64833,43142,27016,684
Flue gas volume flow—total (m3/h) 56,31469,518115,89181,568113,19951,976
Flue gas temperature (°C)input141014181474122712921085
Flue gas composition (vol.%)input
CO0.0000.0000.0000.0000.0000.000
CO221.21721.48020.96819.82619.5535.693
H2O5.6005.3315.80910.73311.07811.202
SO20.0390.0400.0390.0000.0000.000
N271.57671.56371.59767.86467.78074.577
O21.5671.5861.5871.5781.5898.528
Mass flow of dust drifts (kg/h)output194194194194194194
Mass flow of dust drifts—total (kg/h)output194194194194194194
Specific fuel consumption (m3/t) 4.200.000.001.051.0531.5
Flue gas cooling
Flue gas volume flow (m3/h) 56,31469,518115,89181,568113,19951,976
Flue gas temperature (°C) 141014181474122712921085
Flue gas composition DSKDSKDSKDSKDSKDSK
Volumetric cooling air flow (m3/h) 48,00050,00060,00075,00062,00035,000
Flue gas volume flow (m3/h)input56,31469,518115,89181,568113,19951,976
output104,314119,518175,891156,568175,19986,976
Flue gas temperature (°C)input141014181474122712921085
output8138711015686877667
Flue gas composition (vol.%)inputDSKDSKDSKDSKDSKDSK
output
CO000000
CO211.454212.49413.8210.328712.633.40
H2O3.02313.1013.835.59157.166.69
SO20.02130.02350.030.00000.000.00
N274.992274.67474.1273.198271.7576.36
O210.50939.7088.2110.88158.4613.55
Cogeneration I
Flue gas volume flow (m3/h) 11,9518175,891156,568175,199
Flue gas temperature (°C)input 8711015686877
output 800775575570
Theoretical heat output (MW) 3.819.217.4624.02
Flue gas redistribution after cooling (%)Zone 3626267303536
Zone 2383833706564
Flue gas cooling for Zone 2
Flue gas volume flow (m3/h) 39,63945,41758,044109,598113,87955,665
Flue gas temperature (°C) 813800775575570667
Flue gas composition DSKDSKDSKDSKDSKDSK
Volumetric cooling air flow (m3/h) 48,00055,00060,00055,00075,00045,000
Flue gas volume flow (m3/h)input39,63945,41758,044109,598113,87955,665
output87,639100,417118,044164,598188,879100,665
Flue gas temperature (°C)input813800775575570667
output391385409388360381
Flue gas composition (vol.%)inputDSKDSKDSKDSKDSKDSK
output
CO0.0000.0000.0000.0000.0000.000
CO25.1815.6516.7936.8777.6171.881
H2O1.3671.4021.8823.7234.3153.702
SO20.0100.0110.0130.0000.0000.000
N277.18777.04376.60275.13774.62977.538
O216.25515.89314.71114.26313.43816.879
Preheating Zone 3
Mass flow of charge (kg/h)input60,13763,14871,86868,64275,96248,611
output60,13763,14871,86868,64275,96248,611
Temperature of charge (°C)input225225225200200230
output650650650400400500
Flue gas volume flow (m3/h)input64,67574,101117,84746,97061,32031,311
output64,67574,101117,84746,97061,32031,311
Flue gas temperature (°C)input813800775575570667
output527538589378403374
Flue gas composition (vol.%)CO0.0000.0000.0000.0000.0000.000
CO211.45412.49413.81510.32912.6343.402
H2O3.0233.1013.8285.5927.1576.694
SO20.0210.0230.0260.0000.0000.000
N274.99274.67474.12273.19871.75176.357
O210.5099.7088.20910.8818.45813.547
Heat fluxes per charge (kJ/h) 25,51626,05731,20911,99813,61311,796
Useful heat of charge (kJ/h) 25,01326,60131,19811,92213,07711,930
Cogeneration II
Flue gas volume flow (m3/h) ----117,847-
Flue gas temperature (°C)input----589-
output----409-
Theoretical heat output (MW) ----15.47-
Preheating Zone 2
Mass flow of charge (kg/h)input60,13763,14871,86868,64275,96248,611
output60,13763,14871,86868,64275,96248,611
Temperature of charge (°C)input130130130110110140
output225225225200200230
Flue gas volume flow (m3/h)input87,639100,417118,044164,598188,879100,665
output87,639100,417118,044164,598188,879100,665
Flue gas temperature (°C)input391385409388360381
output325322345349321335
Flue gas composition (vol.%)CO0.0000.0000.0000.0000.0000.000
CO25.1815.6516.7936.8777.6171.881
H2O1.3671.4021.8823.7234.3153.702
SO20.0100.0110.0130.0000.0000.000
N277.18777.04376.60275.13774.62977.538
O216.25515.89314.71114.26313.43816.879
Heat fluxes per charge (kJ/h) 499751,1255834805169326429
Useful heat of charge (kJ/h) 484751436000467251133539
Flue gas cooling for Zone 1
Volumetric cooling air flow (m3/h) 40,00040,00040,00040,00035,00023,000
Flue gas volume flow (m3/h)input152,314174,518235,891211,568250,199131,976
output192,314214,518275,891251,568285,199154,976
Flue gas temperature (°C)output305316299283289278
Flue gas composition (vol.%)CO0.0000.0000.0000.0000.0000.000
CO26.2136.9618.8086.4287.7611.909
H2O1.6401.7282.4403.4804.3973.757
SO20.0120.0130.0160.0000.0000.000
N276.82676.49075.89075.38974.54777.517
O215.31014.70812.84514.70313.29616.817
Preheating Zone 1 drying
Mass flow of charge (kg/h)input69,69472,91682,24779,81187,45256,875
output601,3763,14871,86868,64275,96248,611
Temperature of charge (°C)input252525252525
output130130130110110140
Flue gas volume flow (m3/h)input192,314214,518275,891251,568285,199154,976
output204,207226,674288,807265,468299,497165,260
Flue gas temperature (°C)input305316299283289278
output165183188167182132
Flue gas composition (vol.%)input
CO0.0000.0000.0000.0000.0000.000
CO26.2136.9618.8086.4287.761.909
H2O1.6401.7282.4403.4804.403.757
SO20.0120.0130.0160.0000.000.000
N276.82676.59075,89075.38974.5577.517
O215.31014.70812.84514.70313.3018.817
output
CO0.0000.0000.0000.0000.0000.000
CO25.8516.5888.4146.0927.3901.791
H2O7.3686.9986.8038.5348.9619.746
SO20.0110.0120.0160.0000.0000.000
N272.35272.48372.49671.44270.98872.693
O214.41813.92012.27113.93312.66115.771
Mass flow of dust drifts (kJ/h) 194194194194194194
Heat fluxes per charge (kJ/h) 28,22231,89732,16929,44531,28322,953
Useful heat of charge (kJ/h) 29,31730,12932,48032,61133,77025,339
Filter before the chimney
Flue gas volume flow (m3/h)input204,207226,674288,807265,468299,497165,260
output204,207226,674288,807265,468299,497165,260
Flue gas temperature (°C)input163181186165180130
output161179184163178128
Flue gas composition (vol.%)CO0.0000.0000.0000.0000.0000.000
CO25.8516.5888.4146.0927.3901.791
H2O7.3686.9986.8038.5348.9619.746
SO20.0110.0120.0160.0000.0000.000
N272.35272.48372.49671.44270.98872.693
O214.41813.92012.27113.93312.66115.771
Filter efficiency (%) 99.9599.9599.9599.9599.9599.95
Mass flow of dust drifts (kg/h)input194194194194194194
output0.0970.0970.0970.0970.0970.097
Chimney
Flue gas volume flow (m3/h) 204,207226,674288,807265,468299,497165,260
Flue gas temperature in the chimney (°C)input161179184163178128
output155173179158174123
Fuel consumption
Natural gas (kg/h) 66674014805929992146
(m3/h) 9001000200080013502900
Consumption of reducing agent
Coke (kg/h) 858810,83217,330---
Lignite (kg/h) ---20,22827,814-
Total heat (MJ/h) 235,036291,534480,951341,408478,511105,353
Specific fuel consumption (m3/t) 18.5120.5741.1416.4627.7760.87
Specific heat consumption (GJ/t) 4.8355.9979.8947.0239.8442.167
Table 6. Material and energy balance for reduction with coke.
Table 6. Material and energy balance for reduction with coke.
Degree of Reduction50%61%92%
kg/hkg/tkg/hkg/tkg/hkg/t
Dry leach residue53,0941092.2270754,2661116.32282957,6581186.101
Water in the leach residue6381131.2719766522134.16798366930142.5544
Dry matter in coke7042144.87237588882182.722541414,211292.3313
Water in the coke154631.80125323195040.10982615311964.17028
Added water163033.52764336129626.660299463296.773442
Raw pellets total69,6941433.70031872,9161499.98347982,2471691.93
Water leach residue +coke7927163.07322928472174.277809710,049206.7247
Dry matter in raw pellets60,1371237.09944563,1481299.0453771,8681478.432
Added water163033.52764336129626.660299463296.773442
Total water9557196.60087259768200.938109210,378213.4981
Raw pellets total69,6941433.70031872,9161499.98347982,2471691.93
Wet leach residue59,4761223.49904660,7881250.49081264,5871328.655
Wet coke8588176.67362910,832222.832367517,330356.5016
Added water163033.52764336129626.660299463296.773442
Raw pellets total69,6941433.70031872,9161499.98347982,2471691.93
Pellets48,611 48,611 48,611
m3/hm3/tpelletsm3/hm3/tpelletsm3/hm3/tpellets
Gas consumption (NG)90018.51428571100020.57142857200041.14286
Air consumption187,1633850.209098207,8794276.358675264,8755448.863
Produced flue gases204,2074200.836778226,6744663.000971288,8075941.166
MJ/hGJ/tpelletsMJ/hGJ/tpelletsMJ/hGJ/tpellets
Energy consumption235,0364.835031312291,5345.997264592480,9519.893845
Table 7. Material and energy balance for reduction with lignite.
Table 7. Material and energy balance for reduction with lignite.
Degree of Reduction69%92%
kg/hkg/tkg/hkg/t
Dry leach residue53,1901094.19353,6991104.669
Water in the leach residue6393131.50826454132.7673
Dry matter in lignite15,455317.938820,857429.0683
Water in the lignite477398.187596441132.5072
Added water00−1618−33.2941
Raw pellets total79,8111641.82785,8331765.717
Water leach residue + lignite11,165229.68413,138270.2765
Dry matter in raw pellets68,6421412.05975,9621562.657
Added water40.084647−1649−33.9219
Total water11,169229.768611,489236.3546
Raw pellets total79,8111641.82787,4521799.011
Wet leach residue59,5831225.70161,2871260.769
Wet lignite20,228416.126427,814572.1646
Added water 00−1649−33.9219
Raw pellets total79,8111641.82787,4521799.011
Pellets48,611 48,611
m3/hm3/tpelletsm3/hm3/tpellets
Gas consumption (NG)80016.45714135027.77143
Air consumption239,9254935.596268,9205532.058
Produced flue gases265,4685461.054299,4976161.078
MJ/hGJ/tpelletsMJ/hGJ/tpellets
Energy consumption341,4087.02325478,5119.843661
Table 8. Material and energy balance for oxide pellets without reduction.
Table 8. Material and energy balance for oxide pellets without reduction.
kg/hkg/t
Dry leach residue48,6111000
Water in the leach residue5842120.1874
Added water242149.81256
Raw pellets total56,8751170
Water in the leach residue5842120.1874
Dry matter in raw pellets48,6111000
Added water242149.81256
Total water8264170
Raw pellets total56,8751170
Wet leach residue54,4541120.187
Added water242149.81256
Raw pellets total56,8751170
Pellets48,611
m3/hm3/tpellets
Gas consumption (NG)290059.65714
Air consumption134,5152767.168
Produced flue gases165,2603399.639
MJ/hGJ/tpellets
Energy consumption105,3532.167269
Table 9. Overall evaluation of calculation alternatives.
Table 9. Overall evaluation of calculation alternatives.
CokeLigniteOxidized Pellets
Alternative C_50RC_61RC_92RBC_69RBC_92ROX
Degree of reduction%50619269920
Total Fe in pellet%60.361.765.560.462.255.2
Mass flow of raw pelletst/h69,69472,91682,24779,81187,45256,875
Mass flow of pelletst/h48,61148,61148,61148,61148,61148,611
Consumption of raw pellets per 1t of pelletskg/t1433.701499.981691.931641.831799.011170.00
Gas consumptionm3/t18.5120.5741.1416.4627.7759.66
Reducing agent consumptionkg/t176.67222.83356.50416.13572.160.00
Energy consumptionGJ/t4.8355.9979.8947.0239.8442.167
Residence time in RK in reduction zoneMin6060604545-
Required residence time in RK in reduction zoneMin506018070180-
Combustion of preheated air in RK YesYesYesYesYesNo
Need for an afterburner/combustion chamber YesYesYesYesYesYes
Need for mixing chamber I. YesYesYesYesYesYes
Need for flue gas cogeneration I. NoYesYesYesYesNo
Need for flue gas cogeneration II. NoNoYesNoNoNo
Need for mixing chamber II. YesYesYesYesYesYes
Need for mixing chamber III. YesYesYesYesYesYes
Utilization of energy in cogeneration or heat exchangerMW03.79934.6857.46024.020.000
Air consumption for the processm3/t385042765449493655322767
Outgoing flue gas flowm3/t420146635941546161613400
Flue gas temperature before entering the chimney°C165183187.5167182132
Feasibility of the alternative YESYESNoYES *NoYES
* The alternative is feasible only under certain conditions.
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

Varga, A.; Kizek, J.; Rimar, M.; Fedak, M.; Jablonský, G.; Oravec, P.; Bialik, W. Energy Evaluation and Mathematical Modeling of Pellet Production from Metal-Bearing Waste with a Focus on Alternative Applications of Reducing Agents. Processes 2024, 12, 1938. https://doi.org/10.3390/pr12091938

AMA Style

Varga A, Kizek J, Rimar M, Fedak M, Jablonský G, Oravec P, Bialik W. Energy Evaluation and Mathematical Modeling of Pellet Production from Metal-Bearing Waste with a Focus on Alternative Applications of Reducing Agents. Processes. 2024; 12(9):1938. https://doi.org/10.3390/pr12091938

Chicago/Turabian Style

Varga, Augustin, Jan Kizek, Miroslav Rimar, Marcel Fedak, Gustáv Jablonský, Peter Oravec, and Wojciech Bialik. 2024. "Energy Evaluation and Mathematical Modeling of Pellet Production from Metal-Bearing Waste with a Focus on Alternative Applications of Reducing Agents" Processes 12, no. 9: 1938. https://doi.org/10.3390/pr12091938

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