Next Article in Journal
Morphology Modulation of ZnMn2O4 Nanoparticles Deposited In Situ on Carbon Cloth for Supercapacitors
Previous Article in Journal
Assessment of the Thermomechanical Behavior and Microstructure of AA 7075-T6 Aluminum Alloy Lap Joints at Optimal Predicted FSW Process Parameters
Previous Article in Special Issue
Selective Separation and Recovery of Li from Spent LiFePO4 Cathode Materials by Oxidation Roasting Followed by Low-Acid Pressure Leaching
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamic Simulation Model of Copper Side-Blown Smelting Process

1
Zijin Mining Group Co., Ltd., Longyan 361000, China
2
School of Metallurgical Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
*
Author to whom correspondence should be addressed.
Metals 2024, 14(8), 840; https://doi.org/10.3390/met14080840
Submission received: 12 June 2024 / Revised: 21 July 2024 / Accepted: 21 July 2024 / Published: 23 July 2024

Abstract

:
In this study, the thermodynamic simulation model and system of the copper side-blown smelting process were established using the chemical equilibrium constant method, based on the process reaction mechanism, multiphase equilibrium principle, and MetCal software platform (MetCal v7.81). Under typical production conditions, the composition of the product and the distribution behavior of impurity elements were simulated. The results indicate that the average relative error between the calculated mass fractions of major elements such as Cu, S, Fe, SiO2, CaO, MgO, and Al2O3 in copper matte and smelting slag, and the actual production values, is 4.25%. Additionally, the average relative error between the calculated distribution ratios of impurity elements such as Pb, Zn, As, Bi, Mo, Au, and Ag in copper matte and smelting slag, and the actual production data, is 6.74%. Therefore, this model and calculation system accurately reflects the actual production situation of the copper side-blown smelting process well and has potential to predict process output accurately while optimizing process parameters, effectively guiding production practice.

1. Introduction

The side-blown smelting process is a modern and intensified new technology for copper smelting in a bath independently innovated by China [1]. It boasts advantages such as strong adaptability to raw materials, high production efficiency, low investment, low production costs, and environmental friendliness. It is also currently one of the preferred processes for treating complex lead-zinc ore, copper-bearing solid waste and hazardous waste, urban mines, and other difficult-to-treat secondary resources [2,3,4]. More than 20 production lines have been established in China, with an annual capacity of cathode copper accounting for more than 20%, demonstrating a promising application prospect. With the implementation of the national strategies of “carbon peaking and carbon neutrality” and “digitalization”, the digital and intelligent transformation and upgrading of the copper smelting production process has become an important means to further improve quality and efficiency, as well as to conserve energy and reduce emissions.
However, the copper side-blown oxidation smelting process belongs to a complex pyrometallurgical reaction system that involves high temperature, multiple phases, multiple variables, and unstable states [5]. It differs from other bath smelting processes, such as top-blowing [6] and bottom-blowing [7], in terms of the position of air or oxygen-rich gas injection, reaction conditions, control strategies, and technical indicators. Several factors, including raw material degradation, extensive use of recycled materials, unstable batching, and intensified smelting, lead to variable thermodynamic behaviors in the side-blown furnace during raw material decomposition, melting, oxidation, matte formation, and slagging. These behaviors involve phase transformation, reaction evolution, and component migration and distribution. The current manual and empirical approach faces technical challenges in quantitatively and directionally controlling these processes, resulting in frequent abnormal furnace conditions such as slag and flue blockage, difficulties in slag–metal separation, increased copper content in slag, accelerated furnace corrosion, and heightened environmental pollution risks [8]. These issues ultimately impact operational efficiency and performance indicators.
In recent years, to overcome the conditional limitations of traditional experimental methods, improve the efficiency of experimental research, and reduce the cost of experimental research, scholars at home and abroad have applied thermodynamic simulation modeling and analysis techniques to pyrometallurgical research on copper, lead, and other metals, achieving research results that can guide production practices. Guo et al. [9,10], Wang et al. [11,12], Li et al. [13,14], and other researchers focused on pyrometallurgical processes such as bottom-blowing copper smelting furnaces and flash smelting furnaces. Based on the principle of multiphase equilibrium, they employed the methods of minimum free energy and chemical equilibrium constant to conduct process optimization analysis on the technological processes after establishing thermodynamic simulation models. On the other hand, Wang et al. [15], Chen et al. [16], Liu et al. [17,18], and others conducted thermodynamic simulation analysis on lead smelting processes such as flash smelting, bottom-blown smelting, and side-blown smelting by establishing a multiphase equilibrium mathematical model based on the method of chemical equilibrium constants. It can be seen that the main thermodynamic simulation modeling and analysis methods for complex pyrometallurgical multiphase reaction systems include the minimum free energy method [19] and the chemical equilibrium constant method [20]. The former has a simple modeling process but lacks satisfactory convergence in solving, while the latter exhibits good convergence and adaptability to trace element systems. However, most existing studies on the copper side-blown smelting process focus on qualitative analysis of the causes of production problems and empirical summaries of improvement measures based on long-term production practices. There is insufficient research on the mechanisms of phase evolution, component migration and distribution, and phase equilibrium analysis during the smelting process. This has led to the inability to quantitatively determine effective control strategies for the complex side-blown oxidative smelting process of copper concentrates. Currently, the main software platforms that can perform thermodynamic simulation modeling and quantitative calculations for copper pyrometallurgical processes include FactSage, HSC Chemistry, Thermal-Cal, MetCal, and others. Among them, the MetCal software platform (MetCal v7.81) is jointly developed by Jiangxi University of Science and Technology and Jiangxi Maikai Technology Co., Ltd. It features comprehensive thermodynamic data and high efficiency in the secondary development of models. Leveraging this platform, it is possible to quickly construct mathematical models for chemical equilibrium, thermal equilibrium, and mass balance in metallurgical and chemical processes. Currently, the MetCal software platform (MetCal v7.81) has been widely used in thermodynamic simulation analysis research on smelting processes of copper, lead, and other metals [21,22,23,24].
Because of this, based on the principle of multiphase equilibrium in the copper side-blown smelting process, this article establishes a thermodynamic simulation model and calculation system for the copper side-blown smelting process using the method of chemical equilibrium constants and the MetCal software platform (MetCal v7.81). The model simulates and calculates the composition of products and key technical indicators and compares them with industrial production data to verify the accuracy of the model calculations. This is aimed at providing software tools for subsequent predictions of the process output, revealing the behavior patterns of impurity distribution, and optimizing process parameters.

2. Process Mechanism and Mathematical Model Establishment

2.1. Process Reaction Mechanism

The copper side-blown smelting process is usually under the temperature condition of 1200~1350 °C, blowing oxygen-rich air into the tuyeres on both sides of the copper side-blown smelting furnace shown in Figure 1. The raw materials, including copper concentrate, flux, reverts, and pulverized coal, are mixed uniformly in a certain proportion and fed into the furnace through a quantitative belt conveyor and feeder from the top of the furnace. The high-speed oxygen-rich air blown into the furnace vigorously agitates the raw materials and melts, creating favorable gas–liquid–solid three-phase heat and mass transfer kinetic conditions, and the particle size of the raw material to be fed into the furnace is generally less than 50 mm. This allows the raw materials to quickly undergo physicochemical evolution processes such as drying, decomposition, melting, slagging, and matte formation. The produced copper melt enters the electric settling furnace through a siphon and chute to complete the clarification and separation of slag and matte, resulting in copper matte and smelting slag. The flue gas generated during the reaction is cooled in a waste heat boiler, and dust is collected in an electric precipitator and then sent to the sulfuric acid system for acid production.
The main reactions occurring during the side-blown smelting process of copper concentrate include the decomposition of high-valent sulfides, oxidation of sulfides, matte formation, and slagging.
Decomposition of high-valence sulfides:
2 F e S 2 = 2 F e S + S 2
4 CuFeS 2 = 2 Cu 2 S + 4 FeS + S 2
Sulfide oxidation:
2 C uFeS 2 + 4 O 2 = C u 2 S + 2 F e O + 3 S O 2
2 CuS + O 2 = Cu 2 S + SO 2
4 FeS 2 + 11 O 2 = 2 Fe 2 O 3 + 8 SO 2
3 FeS 2 + 8 O 2 = Fe 3 O 4 + 6 SO 2
Matte formation reaction:
FeS + Cu 2 O = FeO + Cu 2 S
x FeS + y Cu 2 S = y Cu 2 S x FeS
Slagging reaction:
2 FeS + 3 O 2 = 2 FeO + 2 SO 2
2 FeO + SiO 2 = 2 FeO SiO 2
Through the above reactions, a copper matte with a higher grade, a smelting slag with a lower copper content, and a suitable slag type, as well as flue gas and dust, can be obtained.

2.2. Modeling Assumption

Based on the reaction mechanism of the copper side-blown smelting process, the products of this process include copper matte, smelting slag, flue gas, and dust. When constructing a multiphase equilibrium model, the equilibrium product phases only include copper matte, smelting slag, and flue gas. Dust is usually considered to be formed by a small amount of unreacted raw materials and fine particle product droplets driven by high-speed airflow in the furnace. Therefore, it is assumed that the composition of dust is consistent with that of a small amount of unreacted raw materials and fine particle product droplets. Based on the above analysis and literature reports, the chemical composition of each product is assumed to be as shown in Table 1. “Other” in the table refers to trace impurity elements, which do not participate in the reaction, but affect the operations involved in modeling the mass conservation relationship.

2.3. Model Construction

The copper side-blown smelting process belongs to a multiphase and multicomponent reaction system, and its thermodynamic model can be constructed using the chemical equilibrium constant method [22], also known as the K-value method. This method involves establishing a mathematical model composed of nonlinear equations based on the total molar number of each element in the system and the independent chemical reactions that occur when the system reaches or approaches equilibrium under constant temperature and pressure conditions. The composition of equilibrium products can then be calculated by solving the model.
For any multiphase multicomponent reaction system, assuming that the number of element types is Ne, and the number of chemical species is Nc, all the components of the system can be determined through reactions among the components. Among them, the number of independent reactions Nb is equal to Nc −Ne, and the independent reactions can be represented by matrix Equation (11).
V j , i A i , k = B j , k
In Equation (11), Vj,i represents the stoichiometric coefficient matrix, Ai,k represents the molecular equation matrix of independent components, and Bj,k represents the molecular equation matrix of dependent components. i, j, and k represent the number of independent components, the number of dependent components, and the number of element types, respectively.
According to the rules of matrix operations, this can be obtained through the operation of Equation (12):
V j , i = B j , k U k , i
In Equation (12), Uk,i is the computable inverse matrix of Aj,k.
For the copper side-blown smelting reaction system, it is assumed that 20% of the inert elements (components) labeled as “Other” are distributed into the copper matte, inert elements refer to trace impurity elements that do not participate in chemical reactions, such as Se, Te, Pt, etc., while 80% of “Other” elements are distributed into the smelting slag. Based on modeling assumptions and phase equilibrium principles, when constructing a multiphase equilibrium model for the copper side-blown smelting process using the chemical equilibrium constant method, the reaction system includes 19 different elements such as Cu, S, Fe, O, Pb, Zn, As, Sb, Bi, Mo, Ag, Au, Si, Ca, Mg, Al, N, H, and C. There are 51 chemical components in the copper matte, smelting slag, and flue gas products involved in the reaction. The molecular equation matrix of independent chemical components consisting of 19 elements is shown in Appendix A Table A1, while the molecular equation matrix of dependent components consisting of the other 32 dependent chemical components is shown in Appendix A Table A2. Therefore, the number of independent reactions in this reaction system is 32. The chemical reactions and their equilibrium constants Kj of the listed independent components are shown in Table 2. The equilibrium constants of independent reactions can be expressed by Equation (13). In Equations (11) and (12), i ranges from 1 to 19, j ranges from 1 to 32, and k ranges from 1 to 19.
K j = exp Δ G b j 0 V j i Δ G a i 0 R T
In Equation (13), R represents the gas constant, T represents the equilibrium temperature of the system, Δ G a i 0 represents the standard Gibbs free energy of formation for the i independent component, and Δ G b i 0 represents the standard Gibbs free energy of formation for the j dependent component.
When the copper side-blown smelting multiphase reaction system reaches equilibrium, the relationship between its independent components and dependent components can be expressed by Equation (14).
Y j = Z m , j γ j K j i γ i X i Z m , i
In Equation (14), Xi represents the molar quantity of the i independent component, γ i represents the activity coefficient of the i independent component, Yj represents the molar quantity of the j dependent component, γ j represents the activity coefficient of the j dependent component, Z m , i represents the molar quantity of the phase belonging to the i independent component, m represents the product phase, and Z m , j represents the molar quantity of the phase belonging to the j dependent component.
For a closed multiphase reaction system, the mass of each element can be calculated using Equation (15) based on the law of conservation of mass.
Q k = i A i , k X i + j B j , k Y j
In Equation (15), Qk represents the molar quantity of the element k.
The total molar quantity Zm of all components in m product phase in Equation (15) can be calculated using Equation (16).
Z m = i m X i + j m Y j
In Equation (16), i ( m ) represents the summation only when the i independent component belongs to the m product phase, and j ( m ) represents the summation only when the j dependent component belongs to the m product phase.
For the copper side-blown smelting multiphase reaction system, assuming that the number of product phases is N p , the system temperature, pressure, and the quantities of each element are given and are close to equilibrium, it can be known from Equations (13)–(15) that the system has a total of N c + N p equations and X i + Y j + Z m variables to be solved, with the number of equations equal to the number of variables to be solved. For the thermodynamic simulation model composed of the nonlinear equations from Equations (13)–(16), the Newton–Raphson algorithm and the algorithm flowchart shown in Figure 2 can be used to solve the model, obtaining the molar quantities of each component in the equilibrium products of each phase in the system.

3. Basic Data and Digital–Analog System

3.1. Raw Materials and Their Composition

The main raw materials used by a domestic copper side-blown smelting production enterprise include mixed copper concentrate, reverts, burning coal, lime powder, quartz sand, air, and industrial oxygen (with an oxygen volume concentration of 95%). Among them, the mixed copper concentrate is configured from copper concentrates from different origins in proportion, the burning coal is configured from coke powder and anthracite in proportion, and air and industrial oxygen are blown into the furnace through primary and secondary air. The elemental composition of the raw materials used by the enterprise from January to December 2018 was obtained through XRF (X-ray fluorescence spectroscopy) analysis. Combined with the analysis results of the main phases of the raw materials through XRD (phase analysis of X-ray diffraction), the phase calculation models of each raw material were constructed separately based on the MetCal platform (MetCal v7.81), and the phase composition of each raw material was calculated and shown in Table 3, Table 4, Table 5, Table 6, Table 7 and Table 8. During the detection processes of XRF and XRD, in order to reduce experimental errors, each sample was tested three times, and the average value was taken as the final result.

3.2. Thermodynamic Basic Data

Based on the relationship between the standard molar reaction enthalpy and temperature as shown in Equation (17) and the relationship between the standard molar reaction entropy and temperature as shown in Equation (18), the Gibbs free energy of the product components in the copper side-blown smelting process is calculated by Equation (19). The standard thermodynamic parameters of the product components are obtained by querying the MetCal software (MetCal v7.81), as detailed in Appendix A Table A3. To eliminate the influence of reaction kinetics, the activity coefficients of some product phases were corrected based on the results of production sampling analysis and concerning the data reported in the literature. Details are shown in Appendix A Table A4. Among them, MQC indicates that the activity of the component is calculated using the modified quasi-chemical solution activity calculation model provided by MetCal v7.81. It is assumed that the flue gas is ideal; therefore, the activity coefficients of the flue gas components are all 1.
Δ H T θ = Δ H 298 θ + 298 T C p d T
Δ S T θ = Δ S 298 θ + 298 T C p T d T
Δ G T θ = Δ H 298 θ T Δ S 298 θ + 298 T C p d T T 298 T C p T d T

3.3. Digital–Analog Computing System

Based on the reaction mechanism of the copper side-blown smelting process and the multiphase equilibrium mathematical model constructed for the copper side-blown smelting process, the thermal equilibrium relationship of this smelting process was established according to the equation shown in Equation (20). The self-developed MetCal software platform (MetCal v7.81) was then utilized to develop a multiphase equilibrium thermodynamic calculation system for the copper side-blown smelting process, as depicted in Figure 3.
i n A Δ H 298 , A i + i n A 298 T i C p A i d T = j n B Δ H 298 , B j + j n B 298 T j C p B j d T + Q Loss
In Equation (20), Ai is the reactant; Ti is the initial temperature of the reactant Ai; Bj is the product; Tj is the temperature of the product Bj; nA is the amount of reactant; nB is the amount of product; H is the enthalpy; Cp is the heat capacity; QLoss is the heat loss.

4. Calculation Examples and Model Verification

4.1. Computational Condition

Using the developed multiphase equilibrium thermodynamic calculation system for the copper side-blown smelting process, the composition of the products in the copper side-blown oxidative smelting production process of a domestic copper smelting enterprise was calculated based on the average operating parameters of the side-blown oxidative smelting section of the enterprise from January to December 2018. The total input of raw materials is 92.6 t/h, and the composition of each raw material is described in Section 3.1. Among them, the mixed copper concentrate is 80.5 t/h, the reverts is 2.01 t/h, the burning coal is 3.40 t/h, the limestone is 4.27 t/h, and the quartz sand is 2.42 t/h. The circulating volume of cooling water is 1100 t/h. The oxygen-to-material ratio is 210 Nm3/t, and the oxygen-rich concentrations of primary air and secondary air are 78% and 26%, respectively. The volume ratio of primary air to secondary air is 4.5. The reverts rate is 2.5%, the burning coal rate is 4.22%, the limestone flux rate is 4.9%, the quartz flux rate is 2.69%, and the smoke dust rate is 1.75%. The temperatures of various smelting products are obtained through thermal balance calculations. It is assumed that the temperature of copper matte is 20 °C lower than that of smelting slag, and the temperature of flue gas and smoke dust is 40 °C higher than that of smelting slag.

4.2. Computation

Under the calculation conditions of the above copper side-blown smelting process, the thermodynamic numerical model and calculation system constructed were used to perform simulation calculations on the process. The chemical composition of copper matte, smelting slag, and dust, as well as the results of thermal balance calculations, are shown in Table 9, Table 10, Table 11 and Table 12.

4.3. Model Verification

In addition to the main elements, the content of impurity elements is one of the important criteria for measuring product quality and is also a key focus in the production process of enterprises, especially the distribution behavior of impurity elements in products such as copper matte and smelting slag. Therefore, to verify the reliability of the constructed model and system, the main element contents and impurity element distribution ratios of products such as copper matte, smelting slag, and smoke dust obtained from simulation calculations were compared with actual production data measured by XRF. The results are listed in Table 13 and Table 14 and Figure 4. The error analysis diagram related to the measurement of production value is shown in Figure 5.
According to the results in Table 13 and Table 14 and Figure 4, the relative errors between the calculated values and production data for the contents of Cu, S, and Fe in copper matte are 1.16%, 2.05%, and 9.63%, respectively. The relative errors for the contents of Cu, S, Fe, SiO2, CaO, MgO, and Al2O3 (pseudo-elements) in smelting slag are 1.27%, 8.08%, 0.28%, 3.45%, 4.70%, 9.12%, and 2.76%, respectively. The relative errors between the calculated values and production data for the distribution ratios of major impurity elements such as Pb, Zn, As, Bi, Mo, Au, and Ag in copper matte and smelting slag are 3.76%, 11.01%, 9.66%, 8.10%, 5.13%, 7.11%, and 2.39%, respectively. The reasons for these errors may include inaccuracies in analysis and detection, as well as inherent errors in the model itself. In order to further improve accuracy and reliability, it is still necessary to optimize the model and conduct more precise detection in the future. Although there are certain errors between the production data and calculated values of the model, these errors are currently within a controllable range. This indicates that the distribution behavior of impurity elements in copper matte and smelting slag obtained from simulation calculations is basically consistent with production practice. It can be seen that the established mathematical model can better reflect the thermodynamic reaction mechanism and characteristics of the copper side-blown smelting process and can be used as an analytical tool for subsequently revealing the material evolution and element distribution behavior patterns of the process.

5. Conclusions

  • Based on the reaction mechanism and characteristics of the copper side-blown smelting process, a multiphase equilibrium thermodynamic calculation model was constructed using the method of chemical equilibrium constants. Based on this, a thermodynamic simulation calculation system was developed, providing a software tool for subsequent thermodynamic simulation analysis of the process;
  • Using the constructed model and calculation system, an example validation of the model was conducted based on the typical operating conditions of the copper side-blown smelting process in a domestic enterprise. The calculated results of the products basically matched the production practice, indicating that the model can basically reflect the reaction characteristics of the copper side-blown process and has the potential to predict the refining production process accurately;
  • Through calculation and comparison, it was found that the calculated values of the main element contents and impurity element distribution ratios in the products of the copper side-blown smelting process had small errors compared with the average measured values of production data. The relative errors of the calculated mass fractions of Cu, S, Fe, SiO2, CaO, MgO, and Al2O3 in copper matte and smelting slag are less than 10%. The relative errors of the distribution ratios of impurity elements such as Pb, Zn, As, Bi, Mo, Au, and Ag in copper matte and smelting slag are less than 11.5%. This indicates that the constructed simulation mathematical model can basically reflect the actual production situation of the copper side-blown smelting process and can be used as an effective tool for subsequent systematic thermodynamic analysis of the process.

Author Contributions

Conceptualization, M.L.; methodology, Y.F. and X.C.; formal analysis, Y.F. and X.C.; writing—original draft preparation, M.L. and Y.F.; writing—review and editing, M.L. and X.C.; funding acquisition, M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Project (52364047) supported by the National Natural Science Foundation of China; Project (20212BAB204026) supported by the Natural Science Foundation of Jiangxi Province of China; Project (2019M662268) supported by the China Postdoctoral Science Foundation; Project (2018KY15) supported by the Postdoctoral Funding Program of Jiangxi Province, China; Project (S202210407006) supported by the Innovation and Entrepreneurship Project for College Students of Jiangxi Province, China.

Data Availability Statement

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

Acknowledgments

We are very grateful to Jindi Huang, Qiang Chen, Fayou He, and Lihua Zhong for their help with this article. We also appreciate the technical support and industrial samples provided by Chifeng Jintong Copper Industry Co., Ltd. and Zijin Mining Group Co., Ltd.

Conflicts of Interest

Author Mingzhou Li was employed by the Zijin Mining Group Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Appendix A

Table A1. Stoichiometry matrix for independent species.
Table A1. Stoichiometry matrix for independent species.
ComponentPhaseCuSFeOPbZnAsSbBiMoAgAuSiCaMgAlNHC
Cu2SMt2100000000000000000
FeSMt0110000000000000000
FeOMt0011000000000000000
Fe3O4Mt0034000000000000000
PbSMt0100100000000000000
ZnSMt0100010000000000000
AsMt0000001000000000000
SbMt0000000100000000000
BiMt0000000010000000000
MoSMt0100000001000000000
Ag2SMt0100000000200000000
AuMt0000000000010000000
SiO2Sl0002000000001000000
CaOSl0001000000000100000
MgOSl0001000000000010000
Al2O3Sl0003000000000002000
N2Gas0000000000000000200
CO2Gas0002000000000000001
H2OGas0001000000000000020
Table A2. Stoichiometry matrix for dependent species.
Table A2. Stoichiometry matrix for dependent species.
ComponentPhaseCuSFeOPbZnAsSbBiMoAgAuSiCaMgAlNHC
Cu2SSl2100000000000000000
Cu2OSl2001000000000000000
FeSSl0110000000000000000
FeOSl0011000000000000000
Fe3O4Sl0034000000000000000
PbOSl0001100000000000000
ZnOSl0001010000000000000
As2O3Sl0003002000000000000
Sb2O3Sl0003000200000000000
Bi2O3Sl0003000020000000000
MoOSl0001000001000000000
AgSl0000000000100000000
AuSl0000000000010000000
SO2Gas0102000000000000000
SO3Gas0103000000000000000
O2Gas0002000000000000000
S2Gas0200000000000000000
PbOGas0001100000000000000
PbSGas0100100000000000000
ZnSGas0100010000000000000
ZnOGas0001010000000000000
ZnGas0000010000000000000
AsOGas0001001000000000000
AsSGas0100001000000000000
As2Gas0000002000000000000
SbOGas0001000100000000000
SbSGas0100000100000000000
SbGas0000000100000000000
BiOGas0001000010000000000
BiSGas0100000010000000000
BiGas0000000010000000000
COGas0001000000000000001
Table A3. Standard thermodynamic parameters of product components.
Table A3. Standard thermodynamic parameters of product components.
ComponentState Δ H 298 Θ /
(kJ·mol−1)
Δ S 298 Θ /
(J·K−1·mol−1)
C p = a + b × 10 3 T + c × 10 5 T 2 + d × 10 6 T 2
abcd
Cu2SLiquid−68.100132.46289.6650.0000.0000.000
Cu2OLiquid−130.22496.40299.9160.0000.0000.000
FeSLiquid−64.63191.20862.5520.0000.0000.000
FeOLiquid−257.27657.59168.2010.0000.0000.000
Fe3O4Liquid−993.334198.385213.3890.0000.0000.000
SiO2Liquid−927.5489.31085.7740.0000.0000.000
CaOLiquid−572.90840.98062.7620.0000.0000.000
MgOLiquid−561.01812.83366.9460.0000.0000.000
Al2O3Liquid−595.56845.145144.8660.0000.0000.000
PbSLiquid−93.14384.12966.9460.0000.0000.000
PbOLiquid−202.24973.37965.0000.0000.0000.000
ZnSLiquid−203.00558.66167.0020.0000.0000.000
ZnOLiquid−309.54247.92060.6690.0000.0000.000
AsLiquid21.56853.28428.8330.0000.0000.000
As2O3Liquid−643.439128.135152.7200.0000.0000.000
SbLiquid17.53162.71231.3810.0000.0000.000
Sb2O3Liquid−675.490143.628156.9040.0000.0000.000
BiLiquid9.27171.98027.1970.0000.0000.000
Bi2O3Liquid−578.024149.814202.0050.0000.0000.000
MoSLiquid−407.113114.979156.9040.0000.0000.000
MoOLiquid358.015302.62038.457−1.797−0.5170.825
AuLiquid0.00047.489−268.634237.1391418.47−52.813
AgLiquid6.39343.22033.4730.0000.0000.000
Ag2SLiquid−32.791142.89393.0020.0000.0000.000
SO2Gas−296.820248.22654.7813.350−24.745−0.241
SO3Gas−395.774256.77877.8344.032−42.617−0.369
N2Gas0.000191.61535.3691.041−41.4650.111
O2Gas0.000205.15434.8601.312−14.1410.163
S2Gas128.603228.16934.6723.286−2.816−0.312
CO2Gas−393.515213.77454.4375.116−43.579−0.806
COGas−110.544197.66529.9325.415−10.814−1.054
H2OGas−241.832188.83731.43814.106−24.952−1.832
PbSGas127.959251.41637.3500.194−2.0960.140
PbOGas68.139240.04841.612−3.526−20.1361.014
ZnSGas204.322236.40427.7137.021251.297−1.105
ZnOGas136.518242.81137.671−0.286−1.9540.735
ZnGas130.403160.99220.898−0.133−0.0670.034
AsOGas43.807230.40843.664−4.280−11.1970.946
AsSGas181.400242.06544.417−4.409−6.8080.916
As2Gas190.711240.88836.7021.152−1.774−0.507
SbOGas−103.502238.35147.257−3.650−40.3240.512
SbSGas190.794249.70146.218−2.657−34.3520.255
SbGas267.181180.2738.9556.15180.063−0.315
BiOGas125.690246.41336.5080.526−3.6630.001
BiSGas176.552257.87838.237−1.090−3.5990.765
BiGas208.742187.01121.189−0.732−0.2030.320
Table A4. Activity coefficients of product components.
Table A4. Activity coefficients of product components.
ComponentProductActivity CoefficientReferences
Cu2SMt1[25,26,27]
FeSMt 0.925 / ( x C u 2 S + 1 ) [25]
FeOMt exp [ 5.1 + 6.2 ( ln x C u 2 S ) + 6.4 ( ln x C u 2 S ) 2 + 2.8 ( ln x C u 2 S ) 3 ] [25]
Fe3O4Mt exp [ 4.96 + 9.9 ( ln x C u 2 S ) + 7.43 ( ln x C u 2 S ) 2 + 2.55 ( ln x C u 2 S ) 3 ] [25]
PbSMt exp [ - 2.716 + 2441 / T + ( 0 . 815 - 3610 / T ) ( 80 [ Pct C u ] mt ) / 100 ] [25]
ZnSMt exp [ - 2.054 + 6917 / T ( 1 . 522 - 1032 / T ) ( 80 [ Pct C u ] mt ) / 100 ] [25]
AsMt exp ( ( 2180 + 3093 ( 80 [ Pct C u ] mt ) / 100 ) / T ) [28]
SbMt exp ( ( 4478 + 3388 ( 80 [ Pct C u ] mt ) / 100 ) / T ) [28]
BiMt exp ( 1900 / T 0.885 ) [29]
MoSMtMQCActivity model
AuMt 10 ( - 3310 / T + 3.15 ) [30,31]
Ag2SMt 10 ( - 425 / T 0.074 + 0.09 x F e S ) [30,31]
Cu2OSl 57.14 x C u 2 O [25]
Cu2SSl exp ( 2.46 + 6.22 N C u 2 S ( m t ) ) [25]
FeSSl70[25]
FeOSl 1.42 x F e O 0.044 [25]
Fe3O4Sl 0.69 + 56.8 x F e 3 O 4 + 5.45 x S i O 2 [25]
SiO2Sl2.1[32]
CaOSl1[32]
MgOSl1[32]
Al2O3Sl1[32]
PbOSl exp ( - 3926 / T ) [30]
ZnOSl exp ( 400 / T ) [33]
As2O3Sl 3.838 exp ( 1523 / T ) P O 2 0.158 [30]
Sb2O3Sl exp ( 1055.66 / T ) [30]
Bi2O3Sl exp ( 1055.66 / T ) [30]
MoOSlMQCActivity model
AuSl480[31,34]
AgSl920[31,35]

References

  1. Hu, Z.Y. Design and Application of Oxygen Enriched Side Blowing for Copper Smelting Process Control System. South. Met. 2023, 9–12+29. [Google Scholar] [CrossRef]
  2. Yuan, J.H. Current Status and Outlook of Side Blowing Furnace. Nonferrous Met. (Extr. Metall.) 2022, 31–35. [Google Scholar] [CrossRef]
  3. Yang, B.; Liu, W.; Jiao, F.; Zhang, L.; Qin, W.; Jiang, S. Numerical Simulation and Application of an Oxygen-Enriched Side-Blown Smelting Furnace for the Treatment of Electroplating Sludge. Sustainability 2023, 15, 10721. [Google Scholar] [CrossRef]
  4. Xie, S.; Zhao, B. Phase Equilibrium Studies of Nonferrous Smelting Slags: A Review. Metals 2024, 14, 278. [Google Scholar] [CrossRef]
  5. Zou, Q.; Hu, J.; Yang, S.; Wang, H.; Deng, G. Investigation of the Splashing Characteristics of Lead Slag in Side-Blown Bath Melting Process. Energies 2023, 16, 1007. [Google Scholar] [CrossRef]
  6. Zhao, B.; Ren, Y.Z.; Jia, W.L.; Zhang, Y.Y.; Zhou, S.W.; Li, B. Regulation of arsenic element trend in copper top-blown smelting process. China Nonferrous Metall. 2024, 53, 88–97. [Google Scholar] [CrossRef]
  7. Wang, W.; Cai, X.; Mu, L.; Lu, T.; Lv, C.; Zhao, H.; Sohn, H.Y. CFD Simulation of the Effects of Mushroom Heads in a Bottom-Blown Copper Smelting Furnace. Metall. Mater. Trans. B-Proc. Metall. Mater. Proc. Sci. 2024, 55, 694–708. [Google Scholar] [CrossRef]
  8. Shi, L.; Wang, Z.Y. Analyzing the Factors Affecting the Concentration of Dilute Acid in the Copper Smelting Process of Oxygen-rich Side Blowing Furance. Non-Ferr. Min. Metall. 2024, 40, 31–33. [Google Scholar]
  9. Guo, X.Y.; Wang, Q.M.; Tian, Q.H.; Zhang, Y.Z. Non-steady multiphase equilbrium process of copperoxygen-enriched bottom blowing bath smelting with gradual changeof oxygen and sulfur potential of different positions in furance. Chin. J. Nonferrous Met. 2015, 25, 1072–1079. [Google Scholar] [CrossRef]
  10. Guo, X.Y.; Wang, Q.M.; Tian, Q.H.; Zhao, B.J. Analysis and optimization of oxygen bottomblowing copper smelting process. Chin. J. Nonferrous Met. 2016, 26, 689–698. [Google Scholar] [CrossRef]
  11. Wang, Q.; Guo, X.; Tian, Q.; Jiang, T.; Chen, M.; Zhao, B. Development and Application of SKSSIM Simulation Software for the Oxygen Bottom Blown Copper Smelting Process. Metals 2017, 7, 431. [Google Scholar] [CrossRef]
  12. Wang, Q.; Guo, X.; Tian, Q.; Jiang, T.; Chen, M.; Zhao, B. Effects of Matte Grade on the Distribution of Minor Elements (Pb, Zn, As, Sb, and Bi) in the Bottom Blown Copper Smelting Process. Metals 2017, 7, 502. [Google Scholar] [CrossRef]
  13. Li, M.; Zhou, J.; Tong, C.; Zhang, W.; Chen, Z.; Wang, J. Thermodynamic Modeling and Optimization of the Copper Flash Converting Process Using the Equilibrium Constant Method. Metall. Mater. Trans. B-Proc. Metall. Mater. Proc. Sci. 2018, 49, 1794–1807. [Google Scholar] [CrossRef]
  14. Li, M.Z.; Zhou, J.M.; Zhang, W.H.; Li, H.S.; Tong, C.R. Thermodynamics analysis of distribution behavior of impurity elements during copper flash converting. Chin. J. Nonferrous Met. 2017, 27, 1951–1959. [Google Scholar] [CrossRef]
  15. Wang, J.L.; Zhang, C.F.; Zhang, W.H. Multi-phase equilibrium model of lead flash smelting process. J. Cent. South Univ. (Sci. Technol.) 2012, 43, 429–434. [Google Scholar]
  16. Chen, L.; Yang, T.Z.; Liu, W.F.; Zhang, D.C.; Bin, S.; Bin, W.D. Distribution of valuable metals in liquid high lead slag during reduction process. Chin. J. Nonferrous Met. 2014, 24, 1056–1062. [Google Scholar] [CrossRef]
  17. Liu, Y.T.; Yang, T.Z.; Li, M.Z. Phase equilibrium model for lead oxygen-enriched side-blown oxidation bath smelting. Chin. J. Nonferrous Met. 2019, 29, 2609–2619. [Google Scholar] [CrossRef]
  18. Liu, Y.T.; Yang, T.Z.; Li, M.Z. Multielement and multiphase equilibrium analysis of lead oxygen-enriched side-blown oxidation smelting. Chin. J. Nonferrous Met. 2020, 30, 1110–1118. [Google Scholar]
  19. Shen, Z.; Li, Y.; Xu, N.; Sun, B.; Du, W.; Xu, M.; Chang, L. Investigation on the Chemical Equilibrium Products for CnHmOlNk Type Fuels Using Equilibrium Constants Database. Fuel 2022, 310, 122325. [Google Scholar] [CrossRef]
  20. Ye, Z.L.; Zhu, Y.F.; Zhang, H.P.; Zhou, S.W.; Li, B.; Shi, Z. A Thermodynamic Study of Copper Oxygen-Riched Smelting Process to Produce High-Grade Matte. J. Kunming Univ. Sci. Technol. (Nat. Sci.) 2022, 47, 54. [Google Scholar] [CrossRef]
  21. Xu, S.C.; Li, M.Z.; Zhao, Z.H.; Zhong, L.H.; He, F.Y.; Huang, J.D. Thermodynamic simulation modeling of copper anode furnace refining process based on MetCal software. Chin. J. Nonferrous Met. 2024, 1–21. [Google Scholar] [CrossRef]
  22. Chen, X.; Li, M.; Liu, F.; Huang, J.; Yang, M. Multi-Phase Equilibrium Model of Oxygen-Enriched Lead Oxidation Smelting Process Based on Chemical Equilibrium Constant Method. Processes 2023, 11, 3043. [Google Scholar] [CrossRef]
  23. Lin, L.; He, F. Numerical Simulation Whole Process of Oxygen-Enriched Side-Blown Smelting and Multi-Lance Top-Blown Converting. Nonferrous Met. (Extr. Metall.) 2023, 103–112. [Google Scholar] [CrossRef]
  24. Xiao, Y.; Lu, T.; Zhou, Y.; Su, Q.; Mu, L.; Wei, T.; Zhao, H.; Liu, F. Computational Fluid Dynamics Study on Enhanced Circulation Flow in a Side-Blown Copper Smelting Furnace. JOM 2021, 73, 2724–2732. [Google Scholar] [CrossRef]
  25. Tan, P.F.; Zhang, C.F. Computer Model of Distribution Behavior of Accessory Elements in Copper Smelting. Acta Metallurgica Sinica 1997, 10, 1094–1100. [Google Scholar] [CrossRef]
  26. Shimpo, R. An Application of Equibibrium Calculations to the Copper Smelting Operation. Available online: https://onemine.org/documents/an-application-of-equibibrium-calculations-to-the-copper-smelting-operation (accessed on 23 April 2024).
  27. Shimpo, R.; Goto, S.; Ogawa, O.; Asakura, I. A Study on the Equilibrium between Copper Matte and Slag. Can. Metall. Q. 1986, 25, 113–121. [Google Scholar] [CrossRef]
  28. Nagamori, M.; Errington, W.J.; Mackey, P.J.; Poggi, D. Thermodynamic Simulation Model of the Isasmelt Process for Copper Matte. Met. Mater. Trans. B 1994, 25, 839–853. [Google Scholar] [CrossRef]
  29. Nagamori, M.; Mackey, P.J.; Tarassoff, P. Copper Solubility in FeO−Fe2O3−SiO2−Al2O3 Slag and Distribution Equilibria of Pb, Bi, Sb and As between Slag and Metallic Copper. Met. Trans. B 1975, 6, 295–301. [Google Scholar] [CrossRef]
  30. Tan, P.; Zhang, C. Computer Model of Copper Smelting Process and Distribution Behaviors of Accessory Elements. J. Cent. South Univ. Technol. 1997, 4, 36–41. [Google Scholar] [CrossRef]
  31. Sinha, S.N.; Sohn, H.Y.; Nagamori, M. Distribution of Gold and Silver between Copper and Matte. Met. Trans. B 1985, 16, 53–59. [Google Scholar] [CrossRef]
  32. Swinbourne, D.R.; Kho, T.S. Computational Thermodynamics Modeling of Minor Element Distributions During Copper Flash Converting. Metall. Mater. Trans. B-Proc. Metall. Mater. Proc. Sci. 2012, 43, 823–829. [Google Scholar] [CrossRef]
  33. Mackey, P.J. The Physical Chemistry of Copper Smelting Slags—A Review. Can. Metall. Q. 1982, 21, 221–260. [Google Scholar] [CrossRef]
  34. Hall, L.D. The Vapor Pressure of Gold and the Activities of Gold in Gold-Copper Solid Solutions. J. Am. Chem. Soc. 1951, 73, 757–760. [Google Scholar] [CrossRef]
  35. Swinbourne, D.R.; Yazawa, A.; Barbante, G.G. Thermodynamic Modeling of Selenide Matte Converting. Met. Mater. Trans. B 1997, 28, 811–819. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of copper side-blown smelting furnace. 1—molten pool; 2—feed opening; 3—furnace stack; 4—flue gas; 5—furnace slag; 6—slag basin; 7—fire bricklayer; 8—blast main; 9—sidewall water jacket; 10—tuyere; 11—copper matte; 12—copper matte pool.
Figure 1. Schematic diagram of copper side-blown smelting furnace. 1—molten pool; 2—feed opening; 3—furnace stack; 4—flue gas; 5—furnace slag; 6—slag basin; 7—fire bricklayer; 8—blast main; 9—sidewall water jacket; 10—tuyere; 11—copper matte; 12—copper matte pool.
Metals 14 00840 g001
Figure 2. Algorithm calculation process.
Figure 2. Algorithm calculation process.
Metals 14 00840 g002
Figure 3. Thermodynamic calculation system of multiphase equilibrium for copper side-blown smelting process.
Figure 3. Thermodynamic calculation system of multiphase equilibrium for copper side-blown smelting process.
Metals 14 00840 g003
Figure 4. The relative error between the model calculation value and the production data. (a) Content of major elements in the product; (b) distribution ratio of impurity elements in the product.
Figure 4. The relative error between the model calculation value and the production data. (a) Content of major elements in the product; (b) distribution ratio of impurity elements in the product.
Metals 14 00840 g004
Figure 5. Measurement error of production value. (a) Copper matte; (b) smelting slag.
Figure 5. Measurement error of production value. (a) Copper matte; (b) smelting slag.
Metals 14 00840 g005
Table 1. The chemical composition of the product.
Table 1. The chemical composition of the product.
PhaseChemical Components
Copper matte (Mt)Cu2S, FeS, FeO, Fe3O4, PbS, ZnS, As, Sb, Bi, MoS, Au, Ag2S, Other1
Smelting slag (Sl)Cu2S, Cu2O, FeS, FeO, Fe3O4, SiO2, CaO, MgO, Al2O3, PbO, ZnO, As2O3, Sb2O3, Bi2O3, MoO, Au, Ag, Other2
Flue gas (Gas)SO2, SO3, N2, O2, S2, PbO, PbS, ZnS, ZnO, Zn, AsO, AsS, As2, SbO, SbS, Sb, BiO, BiS, Bi, CO2, CO, H2O
Dust (Dt)Cu2S, Cu2O, FeS, FeO, Fe3O4, PbS, PbO, ZnS, ZnO, As2O3, As, Sb2O3, Sb, Bi, Bi2O3, MoS, MoO, Au, Ag, Ag2S, SiO2, CaO, MgO, Al2O3, Other3
Table 2. Independent component reaction and equilibrium constant.
Table 2. Independent component reaction and equilibrium constant.
Equilibrium ReactionKjEquilibrium ReactionKj
Cu2S(Mt) + FeO(Mt) = Cu2O(Sl) + FeS(Sl)K12CO(gas) + O2(gas) = 2CO2(gas)K17
2FeS(Mt) + 3O2(gas) = 2FeO(Sl) + 2SO2(gas)K22AsS + 2O2(gas) = As2(gas) + 2SO2(gas)K18
FeS(Mt) = FeS(Sl)K3As2(gas) + O2(gas) = 2AsO(gas)K19
6FeO(Mt) + O2(gas) = 2Fe3O4(Mt)K4SbS(gas) + O2(gas) = Sb(gas) + SO2(gas)K20
2PbS(Mt) + 3O2(gas) = 2PbO(gas) + 2SO2(gas)K52Sb(gas) + O2(gas) = 2SbO(gas)K21
ZnS(Mt) = ZnS(gas)K6BiS(gas) + O2(gas) = Bi(gas) + SO2(gas)K22
2As(Mt) = As2(gas)K72Bi(gas) + O2(gas) = 2BiO(gas)K23
Sb(Mt) = Sb(gas)K82PbS(gas) + 3O2(gas) = 2PbO(gas) + 2SO2(gas)K24
Bi(Mt) = Bi(gas)K9ZnS(gas) + O2(gas) = Zn(gas) + SO2(gas)K25
2MoS(Mt) + 3O2(gas) = 2MoO(Sl) + 2SO2(gas)K10PbO(gas) = PbO(Sl)K26
Au(Mt) = Au(Sl)K114AsO(gas) + O2(gas) = 2As2O3(Sl)K27
Ag2S(Mt) + O2(gas) = 2Ag(Sl) + SO2(gas)K124SbO(gas) + O2(gas) = 2Sb2O3(Sl)K28
2Cu2S(Sl) + 3O2(gas) = 2Cu2O(Sl) + 2SO2(gas)K134BiO(gas) + O2(gas) = 2Bi2O3(Sl)K29
6FeO(Sl) + O2(gas) = 2Fe3O4(Sl)K142Zn(gas) + O2(gas) = 2ZnO(Sl)K30
FeO(Sl) = FeO(Mt)K15S2(gas) + 2O2(gas) = 2SO2(gas)K31
2SO2(gas) + O2(gas) = 2SO3(gas)K162ZnS(gas) + 3O2(gas) = 2ZnO(gas) + 2SO2(gas)K32
Table 3. Elemental content of raw materials (wt.%).
Table 3. Elemental content of raw materials (wt.%).
Raw MaterialCuSFeSiO2CaOMgOAl2O3PbZnAs
Mixed copper concentrate17.99922.31423.58315.2191.2010.8891.2240.3520.7700.101
Reverts21.20010.00031.03718.3402.3361.2404.422---
Burning coal-0.8080.6047.1790.3680.0490.200---
Limestone--0.1944.37050.768-----
Quartz sand--0.46688.7550.9700.5820.970---
Raw MaterialSbBiMoAuAgOCHNOther
Mixed copper concentrate0.0110.0120.1598.29 × 10−43.90 × 10−311.0760.5221.036-3.527
Reverts-----6.858---4.566
Burning coal-----4.92376.4054.1111.3434.011
Limestone-----31.68810.8730.336-1.770
Quartz sand-----3.4180.2080.336-4.296
Table 4. Chemical composition of mixed copper concentrate (wt.%).
Table 4. Chemical composition of mixed copper concentrate (wt.%).
CuFeS2Cu5FeS4FeS2FeSSiO2CaCO3MgCO3Al2O3PbSZnSAs2S3
32.7856.70812.8884.65815.2192.1431.8611.2240.4061.1480.166
Sb2S3Bi2S3MoSAuAg2SH2OFeOFe2O3Cu2OOther
0.0160.0140.2130.0010.0049.2634.6720.3852.7013.525
Table 5. Chemical composition of reverts (wt.%).
Table 5. Chemical composition of reverts (wt.%).
Cu2SCu2OFeSFeOFe3O4SiO2CaOMgOAl2O3Other
26.4110.12312.83425.5524.17618.342.3361.2404.4224.566
Table 6. Chemical composition of burning coal (wt.%).
Table 6. Chemical composition of burning coal (wt.%).
CCH4CO2H2N2H2SFe2O3SiO2
75.3971.20.43.7031.3430.10.8637.179
CaOMgOAl2O3H2OO2SOther
0.3680.0490.20.93.5730.7144.011
Table 7. Chemical composition of limestone (wt.%).
Table 7. Chemical composition of limestone (wt.%).
CaCO3FeOSiO2H2OOther
90.610.254.3731.77
Table 8. Chemical composition of quartz sand (wt.%).
Table 8. Chemical composition of quartz sand (wt.%).
SiO2Fe2O3CaCO3Al2O3MgOH2OOther
88.7550.6661.7310.970.58234.296
Table 9. Chemical composition of copper matte (wt.%).
Table 9. Chemical composition of copper matte (wt.%).
Cu2SFeSFeOFe3O4PbSZnSAsSbBiMoSAuAg2SOther
71.53422.1110.7970.9971.0360.7930.0320.0080.0090.0150.0030.0142.651
Table 10. Chemical composition of smelting slag (wt.%).
Table 10. Chemical composition of smelting slag (wt.%).
Cu2SCu2OFeSFeOFe3O4SiO2CaOMgOAl2O3
1.9170.5520.00132.43713.16433.2407.0341.6552.416
PbOZnOAs2O3Sb2O3Bi2O3MoOAuAgOther
0.1461.2740.1330.0180.0100.3207.44 × 10−53.61 × 10−45.682
Table 11. Chemical composition of dust (wt.%).
Table 11. Chemical composition of dust (wt.%).
Cu2SFeSFeOFe3O4PbSZnSAsSbBi
20.0185.75024.21110.0000.2690.2060.0080.0020.002
MoSAuAg2SCu2OSiO2CaOMgOAl2O3PbO
0.0040.0010.0040.40924.5985.2051.2251.7880.108
ZnOAs2O3Sb2O3Bi2O3MoOAgOther
0.9430.0980.0140.0070.2370.0004.894
Table 12. Heat balance calculation results of copper side-blown smelting process.
Table 12. Heat balance calculation results of copper side-blown smelting process.
Heat IncomeHeat Expense
Heat TypeSuppliesTemp./°CMJ/h%Heat TypeSuppliesTemp./°CMJ/h%
Physical heatMixed copper concentrate250.000.00Physical heatMt117316,873.448.26
Reverts250.000.00Sl119356,541.6427.70
Burning coal250.000.00Gas123380,056.7639.22
Limestone250.000.00St12331708.540.84
Quartz sand250.000.00
Primary oxygen250.000.00
Primary air250.000.00
Secondary oxygen250.000.00
Secondary air250.000.00
Chemical heat 25204,219.43100.00Chemical heat 250.000.00
Exchange heatCooling inlet water39 Exchange heatCooling outlet water4527,612.0613.52
Natural heat dissipation 20018,807.439.21
Hydrocooling 25001.23
Total 204,219.43100.00Total 204,219.43100.00
Table 13. Comparison of main element content of products with production data (wt.%).
Table 13. Comparison of main element content of products with production data (wt.%).
TypePhaseCuSFeSiO2CaOMgOAl2O3
Production dataMt57.79322.42017.0270.178---
Modeling results57.12422.87915.388----
Production dataSl2.0480.42134.64332.1317.3811.8212.351
Modeling results2.0220.38734.73933.2407.0341.6552.416
Table 14. Comparison of impurity element distribution ratio of products with production data.
Table 14. Comparison of impurity element distribution ratio of products with production data.
TypeDxPbZnAsBiMoAuAg
Production dataMt/Sl6.3880.4980.3501.0000.03932.35032.863
Modeling results6.6280.4430.3161.0810.04134.65133.647
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

Li, M.; Feng, Y.; Chen, X. Thermodynamic Simulation Model of Copper Side-Blown Smelting Process. Metals 2024, 14, 840. https://doi.org/10.3390/met14080840

AMA Style

Li M, Feng Y, Chen X. Thermodynamic Simulation Model of Copper Side-Blown Smelting Process. Metals. 2024; 14(8):840. https://doi.org/10.3390/met14080840

Chicago/Turabian Style

Li, Mingzhou, Yuchen Feng, and Xinzhou Chen. 2024. "Thermodynamic Simulation Model of Copper Side-Blown Smelting Process" Metals 14, no. 8: 840. https://doi.org/10.3390/met14080840

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