Next Article in Journal
Quantitative Monitoring Method for Conveyor Belt Deviation Status Based on Attention Guidance
Previous Article in Journal
A Method for Enhancing the Accuracy of Pet Breeds Identification Model in Complex Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Energy and Exergy Analysis of an Improved Hydrogen-Based Direct Reduction Shaft Furnace Process with Waste Heat Recovery

School of Metallurgy, Northeastern University, Shenyang 110819, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2024, 14(16), 6913; https://doi.org/10.3390/app14166913
Submission received: 12 July 2024 / Revised: 3 August 2024 / Accepted: 5 August 2024 / Published: 7 August 2024
(This article belongs to the Special Issue Advanced Processes and Technologies for Sustainable Metallurgy)

Abstract

:

Featured Application

The work in this paper provides some basic theoretical and practical reference value for the energy and exergy analysis, calculation, process optimization, and energy conservation research of the hydrogen-based direct reduction shaft furnace process, which is helpful for promoting its further industrial application.

Abstract

The traditional production mode using coal as the main energy source is not conducive to the sustainable development of the iron and steel industry (ISI). The hydrogen-based direct reduction shaft furnace (HDRSF) process is a feasible technical route for promoting the green development of the ISI. However, there is a lack of comprehensive analysis with respect to the energy utilization and process flow of the HDRSF method. To address these issues, a systemic material–energy–exergy model of HDRSF is established. An improved HDRSF process incorporating waste heat recovery is also proposed, and energy consumption intensity and exergy intensity are used as assessment metrics. This study’s findings indicate that the proposed waste heat recovery can considerably lower gas demand and energy consumption intensity, but exergy intensity has little effect. The reducing gas demand drops from 2083 m3 to 1557 m3, the energy consumption intensity drops from 2.75 × 107 kJ to 1.70 × 107 kJ, and the exergy intensity drops from 1.08 × 107 kJ to 1.05 × 107 kJ when the reducing gas temperature is 900 °C, H2:CO = 1:1; meanwhile, the recovery rate of waste heat reaches 40%. This study can serve as a reference for actual HDRSF process production.

1. Introduction

The ISI plays a crucial role in global economic advancement and industrial growth. Over recent decades, global steel consumption has experienced rapid expansion and is projected to reach 2.2 billion tons by the year 2050 [1]. The ISI is also one of the most energy-intensive and carbon-emitting industries. At present, the average energy consumption per ton of crude steel produced is 20 GJ [2], and carbon dioxide emissions are in the range of 1.8–1.9 tons [3,4]. The ISI contributes approximately 7% to global energy consumption and accounts for 7–9% of global carbon dioxide emissions [5,6]. In the context of global warming and environmental degradation, the high energy consumption and significant carbon emissions associated with traditional production pose substantial challenges to the sustainable development of the ISI [7]. As such, a critical task in the ISI is to accelerate the transformation to green and low-carbon production.
Currently, there exist two primary technical processes for steel production: the blast furnace–basic oxygen furnace (BF-BOF) process and the scrap/direct reduction iron–electric arc furnace (scrap/DRI-EAF) process. Together, these two methods account for over 95% of the world’s total steel production [8,9]. The traditional BF-BOF process is still the most widely used, accounting for about 70% of the world’s total steel production [10]. In the BF-BOF process, about 80% of carbon dioxide emissions originate from the blast furnace ironmaking stage, primarily due to the substantial coal consumption involved [11]. Because coke plays an indispensable structural role in blast furnaces, the research of carbon emission reduction in blast furnace ironmaking has nearly reached a bottleneck [12,13]. The scrap/DRI-EAF process is a typical “non-blast furnace ironmaking” process, where scrap steel or DRI is used as raw material. Compared with the BF-BOF process, the alternative process has shown clear preponderance in energy conservation and carbon emission reduction, leading to rapid development in recent years [14]. However, due to the limited availability of scrap steel resources, relying solely on scrap steel recycling cannot satisfy the rapidly growing global demand for steel [15]. Hence, there is an increasing demand for DRI produced by low-carbon technology [16].
The shaft furnace process plays an important role in DRI production, accounting for more than 72% of global DRI output [17]. The HDRSF approach integrates shaft furnace technology with hydrogen metallurgy. This approach reduces carbon emissions in DRI production at the source by using hydrogen to directly reduce iron ore, a process known as “substitution of carbon with hydrogen”. This promotes cleaner production practices in the ISI [18]. The HDRSF-EAF process, which is a more mature technical route in the practical application of hydrogen metallurgy, can significantly reduce carbon emissions [19]. The flexible mixing ratio of reducing gases such as H2 and CO also renders the HDRSF process more affordable [20]. Owing to the above characteristics, the HDRSF process is expected to be an important route for achieving the green and low-carbon development of the ISI in line with the concept of hydrogen metallurgy [21,22,23].
In recent years, China, the United States, Europe, and many other countries and regions have conducted research on the HDRSF process [24], thereby promoting its development and improvement. Hydrogen is a clean energy source that, when used in the reduction reaction with iron ore, produces only water vapor as a gaseous byproduct. This significantly reduces carbon emissions in steel production. According to calculations by Li et al. [25], compared with the BF-BOF process, which emits 2054.33 kg of carbon per ton of steel, the carbon emissions in the hydrogen-rich shaft furnace–electric arc furnace process based on coal gasification and the pure hydrogen shaft furnace–electric arc furnace process based on renewable electricity and water electrolysis have been reduced to 1121.28 kg and 120.92 kg per ton of steel, respectively. Abhinav et al. [26] established a material and energy model for the production of steel without carbon emissions. The research results reveal that the hydrogen direct reduction iron ore–electric arc furnace process can decrease the carbon emissions of EU steel production exceeding 35%. Rechberg et al. [27] evaluated the carbon emissions of the hydrogen-based direct reduction from the perspective of the carbon flow of electricity in hydrogen production and found that when the carbon emission intensity of electricity was lower than 120 g CO2 kWh−1, the carbon emissions of the HDRSF route were less than that of the natural gas-based direct reduction in the shaft furnace route.
The hydrogen reduction of iron ore involves a highly endothermic reaction that necessitates adequate heat supply to sustain the reduction process at high temperatures. Due to the cost implications of using hydrogen in iron ore reduction, alongside efforts to reduce carbon emissions, several scholars have also focused on significant research into energy efficiency, consumption reduction, and process optimization for the HDRSF method. Chai et al. [28] established a thermal equilibrium model of hydrogen metallurgy and calculated that the optimal reduction temperature in hydrogen metallurgy is 1250–1350 K. The utilization of pure hydrogen for reduction can effectively raise the reaction temperature, thereby reducing overall energy consumption. Qiu et al. [29] established a mass–energy–exergy calculation model of the reduction section of HDRSFs, analyzed the influence of iron ore preheating temperature on the exergy intensity per ton of DRI, and proposed a furnace top gas recovery process. Zhang et al. [30] established a thermal equilibrium model of the reduction section in HDRSFs to explore the energy consumption per ton of DRI at different H2 and CO ratios within a temperature range of 800–1300 °C. After the numerical simulation and calculation of the reduction section of HDRSFs, Ariyan et al. [31] and Shao et al. [32] found that double-row gas injection could enhance the utilization of reducing gas in the shaft furnace and reduce energy consumption. Qiu et al. [33] established a thermal balance model of the reduction section of HDRSFs and an optimization model for the utilization of reducing gas, where the utilization of reducing gas after process optimization increased by 26.7%.
In summary, the HDRSF process represents a significant technical pathway for the low-carbon production of DRI through hydrogen metallurgy technology, crucial for advancing the green and low-carbon development of the ISI. However, this technology is still in its early stage and requires further in-depth analysis and improvements in the energy utilization and process efficiency. Currently, most energy analysis studies of HDRSFs focus on theoretical calculations of hydrogen metallurgical energy consumption, primarily examining the reduction section while neglecting the impact analysis on the cooling section, leading to incomplete assessments of overall shaft furnace energy utilization. Furthermore, the whole HDRSF process is rarely discussed in the existing literature. Given the current high cost of hydrogen acquisition, it is imperative to design and enhance the overall HDRSF process to achieve energy savings, efficiency improvements, and production cost reductions.
Based on the described analysis, in order to further promote the development and improvement of the HDRSF process, the contributions of the present study can be summarized as follows:
  • The principle of DRI production by HDRSFs is introduced. A calculation model for the energy analysis and exergy analysis of HDRSFs is established. Building upon existing research predominantly focused on the reduction section of HDRSFs, the model supplements the analysis by incorporating the cooling section of the shaft furnace. This approach aims to provide a more comprehensive assessment of energy utilization in HDRSFs.
  • An improved process flow of HDRSFs with waste heat recovery is proposed. Under the premise of considering the actual production, the waste heat of top gas and high temperature cooling gas can be recovered and utilized so as to improve the energy utilization efficiency of HDRSFs.
  • Combined with the practical production data of a plant, the influence of waste heat recovery on the reduction gas consumption of HDRSFs is calculated, and the energy flow and exergy flow of the improved HDRSF are analyzed. In addition, two evaluation indicators, namely energy consumption intensity and exergy intensity per ton of DRI, have been established to assess the energy utilization efficiency of the improved HDRSF.
The rest of the article is as follows: Section 2 is an overview of the HDRSF. Section 3 gives an explanation of the model and method description. In Section 4, a process flow of the HDRSF with waste heat recovery is presented. Section 5 includes the analysis and discussion. Section 6 is the conclusion.

2. Description of the HDRSF

In this section, an overview of the production process of DRI with HDRSF and the main chemical reactions in HDRSF are given. The main abbreviations in the following text are explained in Abbreviations section.

2.1. HDRSF Production Process

The shaft furnace is pivotal equipment for producing hydrogen-based DRI [34]. The HDRSF model is shown in Figure 1. Generally, an HDRSF can be divided into three sections from top to bottom: the preheating section, the reduction section, and the cooling section. During normal operation, iron ore enters the shaft furnace from the top and descends due to gravity. The reduction reaction takes place in the reduction section, where the iron ore is exposed to endothermic heat from the preheating section, which raises it to the necessary temperature for reduction. Because there is no obvious boundary between the preheating section and the reduction section, they are collectively referred to as the reduction section for the convenience of modeling and calculation.
Due to economic considerations, it is generally recommended to use the mixture of high-temperature H2 and CO as the reducing gas in HDRSF processing. The heat brought in by the reducing gas is the main heat source of the shaft furnace. The high-temperature mixed reducing gases of H2 and CO flow upward after entering from the middle of the shaft furnace, where there is reduction reaction with the iron ore inside to generate high-temperature DRI. The H2O and CO2 gases generated by the reduction reaction and the unreacted surplus H2 and CO gases are discharged from the top of the shaft furnace. To safeguard the shell of the vertical furnace in the cooling section and prevent the re-oxidation of high-temperature DRI after discharge, it is essential to cool the DRI using gases such as N2 or H2. The cooling gas enters from the lower part of the cooling section and is discharged from the upper part. After the DRI product is cooled to the tap temperature, it is discharged at the bottom of the shaft furnace.

2.2. Main Chemical Reactions

The main chemical reaction in the shaft furnace is the reduction reaction between H2/CO and iron ore. Existing studies have shown that the reduction of iron ore with H2/CO is a multi-stage process [35,36,37]. At T > 570 °C, the reduction step of iron ore with H2/CO is Fe2O3→Fe3O4→FeO→Fe. At T < 570 °C, the reduction step of iron ore with H2/CO is Fe2O3→Fe3O4→Fe. Additionally, according to the balance curve of the reduction of iron oxide with H2/CO, at T > 820 °C, the capacity of H2 to reduce iron oxide is higher than that of CO [38]. In order to fully utilize the advantages of H2 in terms of high reduction rate and low carbon emissions, the reduction temperature in the furnace is usually maintained above 820 °C. At this temperature, the reduction reaction in the HDRSF is shown in Equations (1)–(6) (for simplicity, wüstite is represented by FeO).
Chemical reaction formulas of Fe2O3→Fe with H2:
3 F e 2 O 3 + H 2 = 2 F e 3 O 4 + H 2 O
F e 3 O 4 + H 2 = 3 F e O + H 2 O
F e O + H 2 = F e + H 2 O
Chemical reaction formulas of Fe2O3→Fe with CO:
3 F e 2 O 3 + C O = 2 F e 3 O 4 + C O 2
F e 3 O 4 + C O = 3 F e O + C O 2
F e O + C O = F e + C O 2
Table 1 shows the thermodynamic parameters of the reduction of iron oxide with H2/CO in the step of Fe2O3→Fe3O4→FeO→Fe [24,39], including the enthalpy change, Gibbs free energy, and equilibrium constant.
In actual operation, the reactions in the shaft furnace are reversible, and the equilibrium constant is primarily influenced by temperature. The reduction of iron ore with H2/CO occurs in three stages, following the sequence of the third stage, then the second stage, and finally the first stage after the reducing gas inlets the furnace. The demand for reducing gas varies across different stages of the process. In theory, the equilibrium between gas supply and demand is typically achieved in at most one stage, with excess reducing gas being present in the remaining stages. According to the research of Chai et al. [28], the demand for reducing gas is the greatest in the third stage of the reaction (the reaction in Equations (3) and (6)). The thermodynamic utilization rate of the reducing gas in the third stage can be calculated using Equation (7).
η 3 = x H 2 × K 3 , H 2 1 + K 3 , H 2 + x C O × K 3 , C O 1 + K 3 , C O
where K 3 , H 2 and K 3 , C O are the equilibrium constant of the third-stage reduction reaction with H2 and CO, respectively.
Then, the amount of reducing gas demanded for the chemical reduction reaction to generate DRI that meets the process requirement can be calculated according to Equation (8).
V r e d u c i n g   g a s 22.4 × η 3 = W F e × M D R I 56

3. Shaft Furnace System Modeling

In this section, comprehensive calculation models of the mass balance, energy balance, and exergy balance of the HDRSF are established to analyze and calculate the energy utilization of the hydrogen-based direct reduction process. It should be noted that the calculations in this study are based on the following assumptions: (1) all chemical reactions are in an equilibrium and stable state; (2) gangue does not participate in the reduction reaction; (3) reaction dynamics and material movement are not considered; (4) the interfacial pressure between the cooling section and the reduction section is balanced, and the gases are not mixed; (5) the heat loss at the reduction and cooling sections of the shaft furnace is 15% of its total heat output, respectively [29]; and (6) the standard parameters are 298 K and 0.1 MPa, respectively.

3.1. Mass Balance Model

According to the model of HDRSF in Figure 1, it can be seen that the incoming materials of the shaft furnace include iron ore, reducing gas, and low-temperature cooling gas, while the discharged materials are DRI, top gas, and high-temperature cooling gas. According to the law of conservation of mass, the mass of the materials entering the furnace should be equal to the mass of the materials leaving the furnace. Hence, the mass balance model of the HDRSF is shown in Equation (9).
M i r o n   o r e + M r e d u c i n g   g a s + M c o o l i n g   g a s   i n = M t o p   g a s + M D R I + M c o o l i n g   g a s   o u t
In Equation (9), the mass of DRI depends on the actual production need. According to assumption (5), the mass of discharged high-temperature cooling gas should be the same as the mass of incoming low-temperature cooling gas, and the actual mass of cooling gas needs to be determined according to the DRI yield. In addition to the above three items, the other items in Equation (9) can be calculated by Equations (10)–(12).
M i r o n   o r e = M D R I × W F e T F e × λ F e
M r e d u c i n g   g a s = V r e d u c i n g   g a s × ( ρ H 2 x H 2 + ρ C O x C O )
M t o p   g a s = V t o p   g a s × ( ρ H 2 X H 2 + ρ H 2 O X H 2 O + ρ C O X C O + ρ C O 2 X C O 2 )
In Equations (10)–(12), it can be seen that the mass of iron ore is related to its total iron content, the yield and metallization rate of DRI, and the mass percentage of Fe in DRI. The mass of the reducing gas is related to the volume of the reducing gas and the ratio of H2 and CO in the reducing gas. The top gas includes the surplus reducing gas that is not involved in the reaction, the H2O and CO2 generated by the reduction reaction, and the vapor of water carried in the iron ore. According to the principle of conservation of elements, the extra oxygen in the top gas is equal to the oxygen lost in the iron ore. Based on the mass of oxygen lost in iron ore, the proportion of each component in the top gas in Equation (12) can be calculated by Equations (13)–(17).
X H 2 O = x H 2 O l o s s / 16 + M i r o n   o r e × w H 2 O / 18 V t o p   g a s / 22.4
X C O 2 = x C O O l o s s / 16 V t o p   g a s / 22.4
X H 2 = V r e d u c i n g   g a s × x H 2 / 22.4 x H 2 O l o s s / 16 V t o p   g a s / 22.4
X C O 2 = V r e d u c i n g   g a s × x C O 2 / 22.4 x C O 2 O l o s s / 16 V t o p   g a s / 22.4
The mass of oxygen lost of iron ore in the reduction reaction is the difference between Fe2O3 mass in iron ore and Fe mass in DRI and can be calculated by Equation (17).
O l o s s = M i r o n   o r e × w F e 2 O 3 160 × 16 M D R I × W F e / 56 × 16

3.2. Energy Analysis Model

The reduction of iron ore with hydrogen is a strong endothermic effect, demanding substantial energy input to sustain the reaction. A thorough analysis of energy utilization in the HDRSF process is crucial for identifying recyclable energy sources and minimizing avoidable energy losses. This approach is essential for enhancing the general energy efficiency of the shaft furnace.
The primary focus of existing energy analysis and research on HDRSFs has been on the reduction section, without much attention being paid to the cooling section and overall process flow. The cooling section of the shaft furnace plays a critical role in cooling high-temperature DRI to prevent re-oxidation after discharge and to protect the furnace shell. A comprehensive energy utilization analysis of both the reduction and cooling sections of the HDRSF is depicted in Figure 2, adhering to the principle of energy conservation.
According to the energy budget shown in Figure 2, at the reduction section of the shaft furnace, the energy balance is shown in Equation (18).
Q i r o n   o r e + Q r e d u c i n g   g a s = Q t o p   g a s + Q r e d u c t i o n + Q v a p o r + Q l o s s 1 + Q H D R I
The calculation methods of the items in Equation (18) are shown in Equations (19)–(24).
Q i r o n   o r e = M i r o n   o r e × w i C i × ( T i r o n   o r e 298 )
Q r e d u c i n g   g a s = V r e d u c i n g   g a s 22.4 × ( x H 2 C H 2 + x C O C C O ) × ( T r e d u c i n g   g a s 298 )
Q t o p   g a s = V t o p   g a s 22.4 × ( X H 2 C H 2 + X H 2 O C H 2 O + X C O C C O + X C O 2 C C O 2 ) × ( T t o p   g a s 298 )
Q r e d u c t i o n = M D R I 56 × ( x H 2 Δ H H 2 + x C O Δ H C O )
where Δ H H 2 is the reaction enthalpy per mole of DRI produced in the reduction of iron core with H2; and Δ H C O is the reaction enthalpy per mole of DRI produced in the reduction of iron core with CO.
The water carried in iron ore will be heated up in the furnace and then evaporated into vapor. The heat absorbed by this process is the sum of the heat absorbed when the water heats up to the boiling point of 373 K and the latent heat of the phase transition of evaporation. The calculation method is shown in Equation (23).
Q v a p o r = M i r o n   o r e × w H 2 O × ( C H 2 O ( l ) × ( 373 298 ) + ν H 2 O + C H 2 O ( g ) × ( T t o p   g a s 373 ) )
Q H D R I = M D R I × W i C i × ( T H D R I 298 )
According to the energy budget shown in Figure 2, for the cooling section of the shaft furnace, the energy balance is shown in Equation (25).
Q H D R I + Q c o o l i n g   g a s   i n = Q c o o l i n g   g a s   o u t + Q C D R I + Q l o s s 2
The calculation methods of the items in Equation (25) are shown in Equations (26)–(29).
Q c o o l i n g   g a s   i n = V c o o l i n g   g a s × C c o o l i n g   g a s   i n × ( T c o o l i n g   g a s   i n 298 )
Q c o o l i n g   g a s   o u t = V c o o l i n g   g a s × C c o o l i n g   g a s   o u t × ( T c o o l i n g   g a s   o u t 298 )
Q C D R I = M D R I × W i C i × ( T C D R I 298 )
The specific heat of each substance in the above equations changes with the change in temperature. The relationship between the specific heat of each substance and the temperature change is shown in Equation (29).
C i ( T ) = a + b × 10 3 T + c × 10 5 T 2

3.3. Exergy Analysis Model

The energy conservation equation on the basis of the first law of thermodynamics focuses on the quantitative conservation of energy in various forms during conversion without distinguishing qualitative differences. Exergy, on the other hand, is a physical quantity that measures the quality or usefulness of energy, allowing for an analysis from both quantitative and qualitative perspectives, and it has been widely used in the analysis and evaluation of energy systems [40,41,42]. Figure 3 illustrates the exergy analysis of HDRSFs based on the second law of thermodynamics, which considers energy quality and efficiency.
According to the exergy budget shown in Figure 3, the equilibrium equations of exergy at the reduction and cooling sections of the shaft furnace are shown in Equations (30) and (31), respectively.
E x i r o n   o r e + E x r e d u c i n g   g a s = E x t o p   g a s + E x H D R I + E x l o s s 1
E x H D R I + E x c o o l i n g   g a s   i n = E x c o o l i n g   g a s   o u t + E x C D R I + E x l o s s 2
The calculations of exergy in Equations (30) and (31) are shown in Equations (32)–(37). In the present study, exergy contains physical exergy and chemical exergy, as shown in Equation (32).
E x = E x p h + E x c h
Physical exergy mainly consists of thermal exergy (related to the system temperature and latent heat of the phase transition) and mechanical exergy (related to the system pressure). The physical exergy can be calculated using Equation (33).
E x p h = ( H H 0 ) + T 0 ( S 0 S )
where (HH0) and (SS0) represent enthalpy change and entropy change, respectively; T0 is the reference ambient temperature, which is 298 K in the present study.
In Equation (33), (HH0) can be calculated by Equation (34).
H H 0 = m × C i ( T m ) × ( T T 0 )
where Tm is the average temperature, calculated by (T + T0)/2.
In Equation (33), (SS0) can be calculated using Equations (35) and (36). Equation (35) is employed to compute the entropy change of substances whose properties remain unaffected by pressure variations. Equation (36) is utilized to calculate the entropy change of the substances whose properties are affected by pressure changes, with appropriate corrections being applied [43].
S 0 S = m T T 0 C i ( T ) T d T
S 0 S = m T T 0 C i ( T ) T d T R ln P 0 P
where T is the material temperature; T0 is the reference ambient temperature, which is 298 K in the present study; R is the ideal gas constant, which is 8.314 J/(mol·K); P0 is the reference ambient pressure, which is 0.1 MPa in the present study; and P is the gas pressure.
The chemical exergy of the mixture is related to its components and is not a simple addition of the standard exergy of each component. The chemical exergy of the mixture can be calculated by Equation (37) [44].
E x c h = e i E x i + R T 0 e i ln e i
where ei is the molar percentage of the i-th component in the mixture.

4. An Improved HDRSF Process with Waste Heat Recovery

Under typical conditions, to ensure sufficient reduction reactions in the shaft furnace and achieve the required metallization rate of DRI, the amount of reducing gas introduced into the furnace often exceeds the theoretical demand for the reduction process. This leads to a significant amount of unreacted H2 and CO in the top gas. For economic reasons, it is essential to separate and recover the unreacted reducing gas from the top gas. However, as the top gas is typically above 300 °C, it must be cooled and dehydrated before purification. This approach considers practical factors such as the technical feasibility and cost-effectiveness of gas separation and purification processes [45,46]. In addition to the required endothermic nature of chemical reactions and irreversible heat losses that cannot be recovered, the residual heat carried by the top gas exiting the reduction section and the cooling gas leaving the cooling section can be recovered.
In order to achieve the energy efficiency and consumption reduction of the HDRSF process, an improved HDRSF production process with waste heat recovery is proposed. In the proposed approach, the waste heat of top gas and discharged high-temperature cooling gas is recycled in the HDRSF process, and the surplus reducing gas in the top gas is also purified for recycling. The flow of the improved HDRSF process with waste heat recovery is shown in Figure 4. For the goal of energy efficiency and consumption reduction, the following two main enhancements are implemented:
  • The top gas first enters the heat exchanger to preheat the reducing gas before entering the heating furnace so as to recover and utilize the waste heat of the top gas and reduce the energy consumption during the heating process of the reducing gas in the electric heating furnace.
  • The discharged high-temperature cooling gas is used for preheating the iron ore entering the furnace, thereby realizing recovery of its waste heat and increasing the heat of the iron ore entering the furnace. According to Equation (18), it is evident that increasing the heat content of iron ore entering the furnace can reduce the external energy input required by the shaft furnace, accordingly decreasing the amount of reducing gas needed.
The improved HDRSF process with waste heat recovery operates as follows: External supplementary reducing gas is primarily heated by transferring heat with the top gas. It is then combined with the reducing gas separated from the top gas after undergoing heat transfer, cooling, dehydration, purification, and pressurization processes. Subsequently, the mixed gas is heated to the required process temperature in an electric heating furnace before entering the shaft furnace for iron ore reduction, completing the gas circulation in the reduction section of the shaft furnace. In the cooling section, cooling gas enters from the bottom of the shaft furnace and is discharged from the upper part. The high-temperature cooling gas discharged from the shaft furnace first enters a waste heat recovery device to preheat the iron ore before it enters the furnace. After cooling, purification, and pressurization, the cooling gas returns to the furnace to complete the circulation process. A gas–water heat exchanger is utilized to cool the top gas before dehydration and entry into the furnace, ensuring continuous process operation. The cooling gas after preheating the iron ore is also cooled again by a gas–water heat exchanger. After exiting the gas–water heat exchanger, the cooling water is cooled in a cooling tower and then returns to the heat exchanger to complete the circulation.

5. Results and Discussion

In this section, in terms of the model established in Section 3 and combined with the actual operation parameters of a plant [29,33], the reducing gas demand of a shaft furnace with or without heat recovery is calculated. The energy utilization of the HDRSF process with or without waste heat recovery is analyzed. Two evaluation indicators of energy consumption intensity and exergy intensity are established to estimate the energy utilization of the improved HDRSF process. As the present study focuses on the energy utilization effects of the proposed waste heat recovery improvement on the HDRSF process, the effects of other equipment unrelated to the improvement process are not discussed.

5.1. Parameter Settings

The actual main operating parameters of an HDRSF in a plant are shown in Table 2.
The iron ore composition is shown in Table 3.
Other relevant parameters related to the calculations of the improved process are reasonably set, and the specific values are shown in Table 4. Considering the cost, N2 is used as cooling gas. The heat recovery rate is dependent on the equipment and technology used and is set at 40% in the calculations.
Moreover, the regression coefficients of the relationship between molar heat capacity at level pressure and temperature of the substances are shown in Table A1 in Appendix A [39]. The values of the standard chemical exergy of the components are shown in Table A2 in Appendix A [47].

5.2. Comparison of Reducing Gas Demand in a Shaft Furnace with or without Waste Heat Recovery

The reduction gas demand in the shaft furnace must satisfy both the requirements of the reduction reaction and the heat demand necessary to sustain continuous reduction reactions. This demand is calculated using Equations (8) and (18), respectively, with the theoretical minimum demand for reduction gas being the greater of the two values. The necessity of reducing gas to meet the reduction reaction is affected by the reaction temperature and the composition of the reducing gas. According to Equation (18), the heat source of the shaft furnace originates from the thermal energy carried by both the reduction gas and the iron ore. In the improved process setup, the iron ore is preheated by high-temperature cooling gas before entering the furnace, increasing the thermal energy delivered by the iron ore and consequently reducing the heat brought in by the reduction gas. Therefore, with waste heat recovery, the thermal balance can be achieved with a reduced volume of reduction gas required at the same temperature. Under the temperature condition of 900 °C, the impact of the shaft furnace on the demand for reduction gas with or without waste heat recovery, under various compositions, is illustrated in Figure 5. Obviously, iron ore preheating can effectively reduce the demand for reducing gas in the shaft furnace.

5.3. Mass Balance

The amount of cooling gas in the cooling section can be calculated according to the energy balance of the cooling section in Equation (38).
C c o o l i n g   g a s × V r e d u c i n g   g a s × ρ N 2 × Δ T c o o l i n g   g a s = C D R I × M D R I × Δ T D R I
where ρ N 2 is the nitrogen density; Δ T c o o l i n g   g a s is the temperature difference between the cooling gas in and the cooling gas out; and Δ T D R I is the temperature difference between the DRI out of the reduction section and the DRI out of the furnace.
In the remaining calculations in this section, the reducing gas composition entering the furnace is set as H2:CO = 1:1. According to Figure 5, the reduction gas demand of shaft furnace in the production of per ton of DRI decreases from 2083 m3 to 1557 m3, which decreases by 25.3%. In addition, 1623.05 kg of iron ore and 2277 m3 of cooling gas need to be put into production. The mass balance of HDRSF without or with heat recovery are shown in Table 5 and Table 6, respectively.

5.4. Comparison of the Energy Utilization of the Shaft Furnace System with or without Waste Heat Recovery

5.4.1. Energy Flow Analysis

The control of energy consumption is one of the most pivotal aspects in the HDRSF process, which is directly related to the production cost of DRI. The energy flow of the HDRSF process without or with waste heat recovery are shown in Figure 6 and Figure 7, respectively.
As observed in Figure 6, the sole source of heat that keeps the shaft furnace’s heat consumption stable is the heat brought in by the high-temperature reduction gas. In the absence of waste heat recovery, 2083 m3 of reducing gas is needed to manufacture one ton of DRI at process parameters of 900 °C and H2:CO = 1:1. The energy of reducing gas entering the furnace is 2.62 × 107 kJ. The electric heating furnace provides the energy needed to heat up the reducing gas. Given the electric heating furnace’s conversion efficiency, the amount of external energy required to be input in the absence of waste heat recovery is 2.75 × 107 kJ, which is comparable with the calculation result of Zhang et al. [30]. Furthermore, the primary energy-consuming components of the shaft furnace are the cooling gas heat and the top gas heat, excluding the required energy consumption of reduction reaction heat absorption and furnace body heat dissipation. Figure 7 depicts the HDRSF process’s energy flow following the top gas recovery and high-temperature cooling gas residual heat. Because of the extra energy income from preheating the iron ore, the amount of gas required to produce one ton of DRI decreases to 1557 m3, and the amount of gas energy entering the shaft furnace decreases to 1.96 × 107 kJ. In addition, heating gas in the electric heating furnace uses less energy when the reduction gas is preheated using top gas. Following waste heat recovery, the energy needed is lowered to 1.70 × 107 kJ.

5.4.2. Exergy Flow Analysis

The exergy flow of the HDRSF process without or with waste heat recovery are shown in Figure 8 and Figure 9, respectively.
Overall, the exergy income of the HDRSF process comes from two sources: chemical exergy brought in by external reducing gas and physical exergy brought in by high-temperature reducing gas leaving the electric heating furnace. According to the conservation of elements, whether or not there is waste heat recovery does not change the reducing gas consumed by the reduction reaction in the shaft furnace. Therefore, when there is a reduction gas purification and recovery process, the chemical exergy of the reducing gas consumed in the production process is unchanged, while the chemical exergy is the main source of the overall exergy income of the process. The process’s exergy income has not changed much in comparison with the energy consumption induced by waste heat recovery, and its exergy investment has decreased from 1.08 × 107 kJ in the absence of waste heat recovery to 1.05 × 107 kJ. However, it is essential to note that after waste heat recovery, the shaft furnace’s energy income decreased from 2.48 × 107 kJ to 1.86 × 107 kJ due to a decrease in the demand for reducing gas in the stove. Lowering the volume of gas and top gas also has a positive effect on HDRSF cost reduction.

5.5. Evaluation Indicators

The energy consumption intensity and exergy intensity per ton of DRI are adopted as the evaluation indicators of energy utilization of the improved HDRSF process.
The energy intensity per ton of DRI is calculated according to Equation (39).
E D R I   c o n s u m p t i o n = E I n p u t M D R I
where EInput represents the net energy input of the process to produce MDRI.
The exergy intensity per ton of DRI is calculated according to Equation (40).
E x D R I   c o n s u m p t i o n = E x I n p u t M D R I
where ExInput represents the net exergy input of the process to produce MDRI.
In terms of overall process flow, the energy intensity of HDRSF production per ton of DRI reduced from 2.75 × 107 kJ to 1.70 × 107 kJ, while the exergy intensity decreased from 1.08 × 107 kJ to 1.05 × 107 kJ.

6. Conclusions

The process uses hydrogen metallurgy to produce DRI and reduce fossil fuel consumption in the steel production process, which is an essential technical route to promote the green and sustainable development of the ISI. The study presented in this paper can support the commercialization of HDRSFs and offer some theoretical and practical references for their growth. The main conclusions are as follows:
  • According to the law of conservation of mass and the first and second laws of thermodynamics, a calculation model of the overall material–energy–exergy analysis of HDRSFs has been established. The model comprehensively considers the input and output of the shaft furnace reduction and cooling sections and realizes a more comprehensive analysis and calculation of the energy utilization of HDRSFs.
  • An improved HDRSF process with waste heat recovery is proposed. In the improved process flow, the waste heat recovery of top gas and high-temperature cooling gas is realized by preheating reducing gas and iron ore. In addition, the process of purifying and recovering furnace top gas is rationally designed. The improved HDRSF process has the conditions for practical industrial application.
  • The calculation results show that the waste heat recovery process proposed in this paper can reduce the reducing gas demand for the per-ton DRI produced by the shaft furnace. When the temperature of reducing gas is 900 °C and H2:CO = 1:1, the reducing gas demand per ton of DRI produced by the shaft furnace decreases from 2083 m3 to 1557 m3, which decreases by 25.3%.
  • In terms of the energy analysis and exergy analysis of the HDRSF process, when the recovery rate of waste heat reaches 40%, the energy consumption intensity per ton of DRI decreases from 2.75 × 107 kJ to 1.70 × 107 kJ, showing a remarkable energy-saving effect. The exergy intensity reduces from 1.08 × 107 kJ to 1.05 × 107 kJ, while the exergy input of the shaft furnace reduces from 2.48 × 107 kJ to 1.86 × 107 kJ. In general, the process proposed in this paper can realize the goal of energy conservation and consumption reduction in HDRSF production.

Author Contributions

Conceptualization, Y.J., Z.C. and W.Z.; methodology, Y.J. and Z.C.; software, Y.J.; validation, Y.J., Z.C. and T.J.; formal analysis, Y.J., Z.C. and T.J.; investigation, Y.J. and Z.C.; resources, Y.J. and Z.C.; data curation, Y.J., T.J. and X.L.; writing—original draft preparation, Y.J.; writing—review and editing, Y.J., Z.C., T.J. and X.L.; visualization, Y.J.; supervision, W.Z.; project administration, W.Z.; funding acquisition, W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Key Research and Development Program of China (Grant No. 2017YFA0700300).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

M i r o n   o r e Mass of iron ore, kg w i Proportion of iron ore components, %
M r e d u c i n g   g a s Mass of reducing gas, kg C i Specific heat of each substance, J/(mol·K)
M t o p   g a s Mass of top gas, kg T i r o n   o r e Temperature of iron ore in, K
M D R I Mass of DRI, kg T r e d u c i n g   g a s Temperature of reducing gas in, K
T F e Total iron content of iron ore, % T t o p   g a s Temperature of top gas out, K
λ F e DRI metallization rate, % ν H 2 O Latent heat of water vaporization, kJ/kg
V r e d u c i n g   g a s Volume of reducing gas, m3 W i Proportion of components of DRI, %
ρ H 2 Hydrogen density, kg/m3 T H D R I Temperature of DRI once outside of the reduction section, K
ρ C O Carbon monoxide density, kg/m3 Q c o o l i n g   g a s   i n Heat carried in by cooling gas, kJ
x H 2 Proportion of hydrogen in reducing gas, % V c o o l i n g   g a s Volume of cooling gas, m3
x C O Proportion of carbon monoxide in reducing gas, % T c o o l i n g   g a s   i n Temperature of cooling gas in, K
V t o p   g a s Volume of top gas, m3 Q c o o l i n g   g a s   o u t Heat carried out by cooling gas, kJ
X H 2 Proportion of hydrogen in top gas, % T c o o l i n g   g a s   o u t Temperature of cooling gas out, K
X H 2 O Proportion of water vapor in top gas, % Q C D R I Heat of DRI out of the furnace, kJ
X C O Proportion of carbon monoxide in top gas, % T C D R I Temperature of DRI out of the furnace, K
X C O 2 Proportion of carbon dioxide in top gas, % Q l o s s 2 Heat dissipation of the furnace at the cooling section, kJ
O l o s s Oxygen loss in the iron ore reduction reaction, kg E x i r o n   o r e Exergy carried in by iron ore, kJ
ρ H 2 O Water vapor density, kg/m3 E x r e d u c i n g   g a s Exergy carried in by reducing gas, kJ
ρ C O 2 Carbon dioxide density, kg/m3 E x t o p   g a s Exergy carried out by top gas, kJ
Q i r o n   o r e Heat carried in by iron ore, kJ E x H D R I Exergy at the reduction section carried out by DRI, kJ
Q r e d u c i n g   g a s Heat carried in by reducing gas, kJ E x l o s s 1 Exergy loss at the reduction section, kJ
Q t o p   g a s Heat carried out by top gas, kJ E x c o o l i n g   g a s   i n Exergy carried in by cooling gas, kJ
Q r e d u c t i o n Endothermicity of the reduction reaction, kJ E x c o o l i n g   g a s   o u t Exergy carried out by cooling gas, kJ
Q v a p o r Water evaporates to absorb heat, kJ E x C D R I Exergy carried out by DRI out of the furnace, kJ
Q l o s s 1 Heat dissipation of the furnace at the reduction section, kJ E x l o s s 2 Exergy loss at the cooling section, kJ
Q H D R I Heat of DRI once out of the reaction section, kJ

Appendix A

Table A1. Regression coefficients of the molar heat capacity at level pressure and temperature values for the components.
Table A1. Regression coefficients of the molar heat capacity at level pressure and temperature values for the components.
Chemical Componenta/(J·mol−1·K−1)b/(J·mol−1·K−2)c/(J·mol−1·K)Temperature Range/K
H227.283.260.50298–3000
H2O30.0010.710.34298–2500
CO28.414.10−0.46298–2500
CO244.149.04−8.54298–2500
N227.874.630298–2500
Fe28.18−7.32−2.90298–800
−561.93334.142912.101060–1184
FeO50.798.62−3.31298–1650
Fe2O398.2977.82−14.85298–953
132.687.3601053–1730
CaO49.624.52−6.95298–2888
MgO48.953.14−11.42298–3048
SiO243.8938.79−9.67298–847
58.9110.040847–1696
Al2O3103.8526.27−29.09298–800
120.529.19−48.37800–2327
Table A2. Values of the standard chemical exergy of the components.
Table A2. Values of the standard chemical exergy of the components.
Chemical ComponentValue (kJ/mol)
H2 (g)236.1
H2O (l)0.9
H2O (g)9.5
CO (g)275.1
CO2 (g)19.9
N2 (g)0.72
Fe (s)376.4
FeO (s)127.0
Fe2O3 (s)16.5
CaO (s)110.2
MgO (s)66.8
SiO2 (s)2.8
Al2O3 (s)200.4

References

  1. Hasanbeigi, A.; Arens, M.; Price, L. Alternative emerging ironmaking technologies for energy-efficiency and carbon dioxide emissions reduction: A technical review. Renew. Sustain. Energy Rev. 2014, 33, 645–658. [Google Scholar] [CrossRef]
  2. Sun, W.; Wang, Q.; Zhou, Y.; Wu, J. Material and energy flows of the iron and steel industry: Status quo, challenges and perspectives. Appl. Energy 2020, 268, 114946. [Google Scholar] [CrossRef]
  3. Ledari, M.B.; Khajehpour, H.; Akbarnavasi, H.; Edalati, S. Greening steel industry by hydrogen: Lessons learned for the developing world. Int. J. Hydrogen Energy 2023, 48, 36623–36649. [Google Scholar] [CrossRef]
  4. Sun, C.; Wang, J.; Zhou, M.; Hong, L.; Ai, L.; Wen, L. Process path for reducing carbon emissions from steel industry—Combined electrification and hydrogen reduction. Processes 2024, 12, 108. [Google Scholar] [CrossRef]
  5. Kim, J.; Sovacool, B.K.; Bazilian, M.; Griffiths, S.; Lee, J.; Yang, M.; Lee, J. Decarbonizing the iron and steel industry: A systematic review of sociotechnical systems, technological innovations, and policy options. Energy Res. Soc. Sci. 2022, 89, 102565. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Yu, L.; Cui, K.; Wang, H.; Fu, T. Carbon capture and storage technology by steel-making slags: Recent progress and future challenges. Chem. Eng. J. 2023, 455, 140552. [Google Scholar] [CrossRef]
  7. Liu, W.; Zuo, H.; Wang, J.; Xue, Q.; Ren, B.; Yang, F. The production and application of hydrogen in steel industry. Int. J. Hydrogen Energy 2021, 46, 10548–10569. [Google Scholar] [CrossRef]
  8. Rahmatmand, B.; Tahmasebi, A.; Lomas, H.; Honeyands, T.; Koshy, P.; Hockings, K.; Jayasekara, A. A technical review on coke rate and quality in low-carbon blast furnace ironmaking. Fuel 2023, 336, 127077. [Google Scholar] [CrossRef]
  9. Kildahl, H.; Wang, L.; Tong, L.; Ding, Y. Cost effective decarbonisation of blast furnace–basic oxygen furnace steel production through thermochemical sector coupling. J. Clean. Prod. 2023, 389, 135963. [Google Scholar] [CrossRef]
  10. Andrade, C.; Desport, L.; Selosse, S. Net-negative emission opportunities for the iron and steel industry on a global scale. Appl. Energy 2024, 358, 122566. [Google Scholar] [CrossRef]
  11. Sun, Y.; Tian, S.; Ciais, P.; Zeng, Z.; Meng, J.; Zhang, Z. Decarbonising the iron and steel sector for a 2 °C target using inherent waste streams. Nat. Commun. 2022, 13, 297. [Google Scholar] [CrossRef] [PubMed]
  12. Yilmaz, C.; Wendelstorf, J.; Turek, T. Modeling and simulation of hydrogen injection into a blast furnace to reduce carbon dioxide emissions. J. Clean. Prod. 2017, 154, 488–501. [Google Scholar] [CrossRef]
  13. Wang, H.; Liu, F.; Zeng, H.; Liao, J.; Wang, J.; Lai, C. Distribution behavior of impurities during the hydrogen reduction ironmaking process. Metals 2024, 14, 718. [Google Scholar] [CrossRef]
  14. An, R.; Yu, B.; Li, R.; Wei, Y. Potential of energy savings and CO2 emission reduction in China’s iron and steel industry. Appl. Energy 2018, 226, 862–880. [Google Scholar] [CrossRef]
  15. Quader, M.A.; Ahmed, S.; Ghazilla, R.A.R.; Ahmed, S.; Dahari, M. A comprehensive review on energy efficient CO2 breakthrough technologies for sustainable green iron and steel manufacturing. Renew. Sustain. Energy Rev. 2015, 50, 594–614. [Google Scholar] [CrossRef]
  16. Ling, J.; Yang, H.; Tian, G.; Cheng, J.; Wang, X.; Yu, X. Direct reduction of iron to facilitate net zero emissions in the steel industry: A review of research progress at different scales. J. Clean. Prod. 2024, 441, 140933. [Google Scholar] [CrossRef]
  17. Shahabuddin, M.; Brooks, G.; Rhamdhani, M.A. Decarbonisation and hydrogen integration of steel industries: Recent development, challenges and technoeconomic analysis. J. Clean. Prod. 2023, 395, 136391. [Google Scholar] [CrossRef]
  18. Li, F.; Chu, M.; Tang, J.; Liu, Z.; Zhou, Y.; Wang, J. Exergy analysis of hydrogen-reduction based steel production with coal gasification-shaft furnace-electric furnace process. Int. J. Hydrogen Energy 2021, 46, 12771–12783. [Google Scholar] [CrossRef]
  19. Tang, J.; Chu, M.; Li, F.; Feng, C.; Liu, Z.; Zhou, Y. Development and progress on hydrogen metallurgy. Int. J. Miner. Metall. Mater. 2020, 27, 713–723. [Google Scholar] [CrossRef]
  20. Wang, R.; Zhao, Y.; Babich, A.; Senk, D.; Fan, X. Hydrogen direct reduction (H-DR) in steel industry—An overview of challenges and opportunities. J. Clean. Prod. 2021, 329, 129797. [Google Scholar] [CrossRef]
  21. Li, Z.; Qi, Z.; Zhang, L.; Guo, M.; Liang, D.; Dong, Q. Numerical simulation of H2-intensive shaft furnace direct reduction process. J. Clean. Prod. 2023, 409, 137059. [Google Scholar] [CrossRef]
  22. Wang, X.; Wang, T.; Chu, M. Progress and Key Problems of Hydrogen Metallurgy, 1st ed.; Chemical Industry Press: Beijing, China, 2023; pp. 46–70. [Google Scholar]
  23. Shao, L.; Xu, J.; Saxén, H.; Zou, Z. A numerical study on process intensification of hydrogen reduction of iron oxide pellets in a shaft furnace. Fuel 2023, 348, 128375. [Google Scholar] [CrossRef]
  24. Zhang, J.; Li, K.; Liu, Z.; Yang, T. Primary Exploration of Hydrogen Metallurgy, 1st ed.; Springer: Singapore, 2024; pp. 117–171. [Google Scholar]
  25. Li, F.; Chu, M.; Tang, J.; Liu, Z.; Zhao, Z.; Liu, P.; Yan, R. Quantifying the energy saving potential and environmental benefit of hydrogen-based steelmaking process: Status and future prospect. Appl. Therm. Eng. 2022, 211, 118489. [Google Scholar] [CrossRef]
  26. Bhaskar, A.; Assadi, M.; Somehsaraei, H.N. Decarbonization of the iron and steel industry with direct reduction of iron ore with green hydrogen. Energies 2020, 13, 758. [Google Scholar] [CrossRef]
  27. Rechberger, K.; Spanlang, A.; Sasiain Conde, A.; Wolfmeir, H.; Harris, C. Green hydrogen-based direct reduction for low-carbon steelmaking. Steel Res. Int. 2020, 91, 2000110. [Google Scholar] [CrossRef]
  28. Chai, X.; Yue, Q.; Zhang, Y.; Wang, Q. Analysis of energy consumption and its influencing factors in hydrogen metallurgy process. Steel Res. Int. 2022, 93, 2100730. [Google Scholar] [CrossRef]
  29. Qiu, Z.; Du, T.; Yue, Q.; Na, H.; Sun, J.; Yuan, Y.; Che, Z.; Wang, Y.; Li, Y. A multi-parameters evaluation on exergy for hydrogen metallurgy. Energy 2023, 281, 128279. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Yue, Q.; Chai, X.; Wang, Q.; Lu, Y.; Ji, W. Analysis of process parameters on energy utilization and environmental impact of hydrogen metallurgy. J. Clean. Prod. 2022, 361, 132289. [Google Scholar] [CrossRef]
  31. Ghadi, A.Z.; Valipour, M.S.; Biglari, M. CFD simulation of two-phase gas-particle flow in the Midrex shaft furnace: The effect of twin gas injection system on the performance of the reactor. Int. J. Hydrogen Energy 2017, 42, 103–118. [Google Scholar] [CrossRef]
  32. Shao, L.; Zhang, X.; Zhao, C.; Qu, Y.; Saxén, H.; Zou, Z. Computational analysis of hydrogen reduction of iron oxide pellets in a shaft furnace process. Renew. Energy 2021, 179, 1537–1547. [Google Scholar] [CrossRef]
  33. Qiu, Z.; Yue, Q.; Yan, T.; Wang, Q.; Sun, J.; Yuan, Y.; Che, Z.; Wang, Y.; Du, T. Gas utilization optimization and exergy analysis of hydrogen metallurgical shaft furnace. Energy 2023, 263, 125847. [Google Scholar] [CrossRef]
  34. Patisson, F.; Mirgaux, O. Hydrogen Ironmaking: How It Works. Metals 2020, 10, 922. [Google Scholar] [CrossRef]
  35. Chen, Z.; Dang, J.; Hu, X.; Yan, H. Reduction kinetics of hematite powder in hydrogen atmosphere at moderate temperatures. Metals 2018, 8, 751. [Google Scholar] [CrossRef]
  36. Elzohiery, M.; Sohn, H.Y.; Mohassab, Y. Kinetics of hydrogen reduction of magnetite concentrate particles in solid state relevant to flash ironmaking. Steel Res. Int. 2017, 88, 290–303. [Google Scholar] [CrossRef]
  37. Zakeri, A.; Coley, K.S.; Tafaghodi, L. Hydrogen-based direct reduction of iron oxides: A review on the influence of impurities. Sustainability 2023, 15, 13047. [Google Scholar] [CrossRef]
  38. Sun, M.; Pang, K.; Barati, M.; Meng, X. Hydrogen-based reduction technologies in low-carbon sustainable ironmaking and steelmaking: A review. J. Sustain. Metall. 2024, 10, 10–25. [Google Scholar] [CrossRef]
  39. Ren, S.; Bai, M.; Long, H.; Xu, K. Technology and Simulation of Direct Reduction in Gas-Based Shaft Furnace, 1st ed.; Metallurgical Industry Press: Beijing, China, 2018; pp. 10–14. [Google Scholar]
  40. Lucia, U.; Grisolia, G. Exergy inefficiency: An indicator for sustainable development analysis. Energy Rep. 2019, 5, 62–69. [Google Scholar] [CrossRef]
  41. Lucia, U. Entropy and exergy in irreversible renewable energy systems. Renew. Sustain. Energy Rev. 2013, 20, 559–564. [Google Scholar] [CrossRef]
  42. Hu, Z.; He, D.; Zhao, H. Multi-objective optimization of energy distribution in steel enterprises considering both exergy efficiency and energy cost. Energy 2023, 263, 125623. [Google Scholar] [CrossRef]
  43. Jiang, T.; Zhang, W.; Liu, S. Performance evaluation of a full-scale fused magnesia furnace for MgO production based on energy and exergy analysis. Energies 2021, 15, 214. [Google Scholar] [CrossRef]
  44. Tang, X. Heat Energy Conversion and Utilization, 2nd ed.; Metallurgical Industry Press: Beijing, China, 2004; p. 37. [Google Scholar]
  45. Luberti, M.; Ahn, H. Review of Polybed pressure swing adsorption for hydrogen purification. Int. J. Hydrogen Energy 2022, 47, 10911–10933. [Google Scholar] [CrossRef]
  46. Karimi, M.; Shirzad, M.; Silva, J.A.C.; Rodrigues, A.E. Carbon dioxide separation and capture by adsorption: A review. Environ. Chem. Lett. 2023, 21, 2041–2084. [Google Scholar] [CrossRef] [PubMed]
  47. Morris, D.R.; Szargut, J. Standard chemical exergy of some elements and compounds on the planet earth. Energy 1986, 11, 733–755. [Google Scholar] [CrossRef]
Figure 1. Model of the HDRSF.
Figure 1. Model of the HDRSF.
Applsci 14 06913 g001
Figure 2. Analysis of the energy balance of the HDRSF.
Figure 2. Analysis of the energy balance of the HDRSF.
Applsci 14 06913 g002
Figure 3. Analysis of the exergy balance of HDRSFs.
Figure 3. Analysis of the exergy balance of HDRSFs.
Applsci 14 06913 g003
Figure 4. The flow of the improved HDRSF process with waste heat recovery.
Figure 4. The flow of the improved HDRSF process with waste heat recovery.
Applsci 14 06913 g004
Figure 5. Comparison of reducing gas requirements of different components in a shaft furnace with or without ore preheating.
Figure 5. Comparison of reducing gas requirements of different components in a shaft furnace with or without ore preheating.
Applsci 14 06913 g005
Figure 6. Energy flow analysis of the HDRSF process without waste heat recovery.
Figure 6. Energy flow analysis of the HDRSF process without waste heat recovery.
Applsci 14 06913 g006
Figure 7. Energy flow analysis of the HDRSF process with waste heat recovery.
Figure 7. Energy flow analysis of the HDRSF process with waste heat recovery.
Applsci 14 06913 g007
Figure 8. Exergy flow analysis of the HDRSF process without waste heat recovery.
Figure 8. Exergy flow analysis of the HDRSF process without waste heat recovery.
Applsci 14 06913 g008
Figure 9. Exergy flow analysis of the HDRSF process with waste heat recovery.
Figure 9. Exergy flow analysis of the HDRSF process with waste heat recovery.
Applsci 14 06913 g009
Table 1. Thermodynamic parameters of reduction of iron oxide.
Table 1. Thermodynamic parameters of reduction of iron oxide.
Reaction Formula Δ H 298 / J mol 1 Δ r G m / J mol 1 lg K
3 F e 2 O 3 + H 2 = 2 F e 3 O 4 + H 2 O −100,070−15,547 − 74.40 T 812 T + 3.886
F e 3 O 4 + H 2 = 3 F e O + H 2 O 60,45071,940 − 73.62 T 3757 T + 3.845
F e O + H 2 = F e + H 2 O 30,23023,430 − 16.16 T 1224 T + 0.844
3 F e 2 O 3 + C O = 2 F e 3 O 4 + C O 2 −23,410−52,131 − 41.00 T 2723 T + 2.141
F e 3 O 4 + C O = 3 F e O + C O 2 19,29035,380 − 40.16 T 1848 T + 2.097
FeO + CO = Fe + C O 2 −10,930−13,175 + 17.24 T 688 T 0.900
Table 2. Main operating parameters of the HDRSF.
Table 2. Main operating parameters of the HDRSF.
ParameterValue
Temperature of top gas out/K623
Temperature of reducing gas in/K1173
Temperature of DRI out of reduction section/K1123
DRI metallization rate/%92
Table 3. Iron ore composition/%.
Table 3. Iron ore composition/%.
TFeFeOCaOSiO2MgOAl2O3H2O
66.790.131.241.690.140.271.00
Table 4. Other relevant parameter settings.
Table 4. Other relevant parameter settings.
ParameterValue
Temperature of cooling gas (N2) in/K323
Temperature of cooling gas (N2) out/K573
Temperature of DRI out of furnace/K353
Recovery rate of waste heat/%40
Electric heating furnace efficiency/%95
Table 5. Mass balance of the HDRSF without waste heat recovery.
Table 5. Mass balance of the HDRSF without waste heat recovery.
InputOutput
ItemMass/kgPercentage/%ItemMass/kgPercentage/%
Iron ore1623.0527.68DRI100017.05
Reducing gas1394.723.78Top gas1963.5433.48
Cooling gas in2846.2548.54Cooling gas out2846.2548.54
Gangue54.210.93
Total5864100Total5864100
Table 6. Mass balance of the HDRSF with waste heat recovery.
Table 6. Mass balance of the HDRSF with waste heat recovery.
InputOutput
ItemMass/kgPercentage/%ItemMass/kgPercentage/%
Iron ore1623.0529.45DRI100018.14
Reducing gas1042.218.91Top gas1611.0429.23
Cooling gas in2846.2551.64Cooling gas out2846.2551.64
Gangue54.210.99
Total5511.5100Total5511.5100
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

Ji, Y.; Chi, Z.; Jiang, T.; Liu, X.; Zhang, W. Energy and Exergy Analysis of an Improved Hydrogen-Based Direct Reduction Shaft Furnace Process with Waste Heat Recovery. Appl. Sci. 2024, 14, 6913. https://doi.org/10.3390/app14166913

AMA Style

Ji Y, Chi Z, Jiang T, Liu X, Zhang W. Energy and Exergy Analysis of an Improved Hydrogen-Based Direct Reduction Shaft Furnace Process with Waste Heat Recovery. Applied Sciences. 2024; 14(16):6913. https://doi.org/10.3390/app14166913

Chicago/Turabian Style

Ji, Yuzhang, Zhongyuan Chi, Tianchi Jiang, Xin Liu, and Weijun Zhang. 2024. "Energy and Exergy Analysis of an Improved Hydrogen-Based Direct Reduction Shaft Furnace Process with Waste Heat Recovery" Applied Sciences 14, no. 16: 6913. https://doi.org/10.3390/app14166913

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