Next Article in Journal
Storage Duration and Added Docosahexaenoic Acid Modify the Rates of Esterified and Free Oxylipin Formation in Infant Milk Formula
Next Article in Special Issue
Online Dynamic Optimization of Multi-Rate Processes with the Case of a Fluid Catalytic Cracking Unit
Previous Article in Journal
Polyethylene Composite Particles as Novel Water Plugging Agent for High-Temperature and High-Salinity Reservoirs
Previous Article in Special Issue
A Simple Strategy for the Simultaneous Determination of Dopamine, Uric Acid, L-Tryptophan and Theophylline Based on a Carbon Nano-Onions Modified Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Phase Equilibrium Model of Oxygen-Enriched Lead Oxidation Smelting Process Based on Chemical Equilibrium Constant Method

Faculty of Materials Metallurgy and Chemistry, Jiangxi University of Science and Technology, Kejia Road No. 156, Ganzhou 341000, China
*
Authors to whom correspondence should be addressed.
Submission received: 7 September 2023 / Revised: 15 October 2023 / Accepted: 20 October 2023 / Published: 23 October 2023
(This article belongs to the Special Issue Chemical Process Modelling and Simulation)

Abstract

:
With the increasingly complicated sources of lead smelting materials, it is becoming more difficult to optimize process parameters during the bottom-blowing lead oxidation smelting process. Building a bottom-blowing lead smelting thermodynamic model has significant importance for the green production of the lead smelting process. In this study, we built a multi-phase equilibrium thermodynamic model and simulation system for the oxygen-enriched bottom-blowing lead oxidation smelting process using the MetCal software platform (MetCal v7.81) according to the chemical equilibrium constant method. The equilibrium products composition and important technical indicators were calculated under factory operating conditions. Compared with the industrial data, the calculation results demonstrated that the average relative error of the calculation value of the mass fraction in the crude lead, lead-rich slag, and dust was 3.76%. The average relative error of important technical indicators, including dust rate, crude lead yield, lead-rich slag temperature, slag iron–silica ratio (RFe/SiO2), and slag calcium–silica ratio (RCaO/SiO2), was 6.39%. As a result, the developed modeling and simulation system was able to reflect the current state of the oxygen-enriched bottom-blowing lead smelting. It also demonstrated the potential to enhance the smelting process and optimize the process parameters. Therefore, it is expected to provide a useful tool for thermodynamic analysis.

1. Introduction

The main production process for primary lead is pyrometallurgy [1]. The mainstream processes of lead smelting include bottom-blowing (SKS) [2], Queneau–Schuhmann–Lurgi (QSL) [3], oxygen-enriched side-blowing [4], Ausmelt [5], Isasmelt [6,7], Kivcet [8], and Outokumpu. Among them, the bottom-blowing smelting process, also known as the SKS technique, was developed by China Enfi Engineering Technology Co., Ltd. in 1998. Because of its advantages of wide adaptability of raw materials, low energy consumption, and environmental friendliness, this process occupies an important position in China’s lead smelting process. As a result of the complicated sources of raw materials and variable compositions [9], it is more challenging to optimize process parameters for the bottom-blowing lead oxidation smelting process. How to optimize process parameters is a crucial and challenging issue in enhancing the bottom-blowing lead smelting process.
These parameters are closely related to the thermodynamics of the bottom-blowing lead smelting process. Oxygen-enriched bottom-blowing smelting is a high-temperature, multi-phase, and multicomponent complex system. Conventional static experiments cannot be used to carry out systematic thermodynamic analysis [10]. Due to the vigorous agitation of oxygen-enriched air on the high-temperature molten pool, the mass and heat transfer processes in the reactor are intensified, which brings the reaction close to thermodynamic equilibrium rapidly [8,11,12,13,14]. Therefore, using a computer to build thermodynamic models to study the fundamental theory of the lead smelting process has increasingly attracted the attention of researchers. Constructing thermodynamic models is mainly based on the minimum Gibbs free energy method [11] or the chemical equilibrium constant method [15,16]. Chen et al. [17] established thermodynamic models for the oxygen-enriched bottom-blowing lead oxidation smelting process with the minimum Gibbs free energy method. Tan et al. [18] and Zhang et al. [19] constructed thermodynamic models for the QSL process based on the chemical equilibrium constant method. However, the existing QSL thermodynamic models did not consider As, Sb, Bi and other impurity elements. They were unable to adapt to the complicated raw materials [9]. Therefore, we used the chemical equilibrium constant method to construct a thermodynamic model, taking into account the impurity elements such as As, Sb, Bi, Mg, and Al.
In the lead smelting research, there were many papers that simulated lead smelting by constructing thermodynamic models [18,19,20], but few of them further developed thermodynamic models into visual operation interfaces. The learning and usage of thermodynamic models became more challenging without visual operation interfaces. In copper smelting, Wang et al. [12] constructed a thermodynamic model for bottom-blown copper smelting and further developed the SKSSIM simulation software (SKSSIM v1.0) with C# program language. The software significantly dropped the difficulty of model usage and improved the efficiency of optimizing process parameters. However, with the increasing importance of the bottom-blowing lead smelting process in China, there is still no researcher who has constructed a thermodynamic model with a visual operating interface for the bottom-blowing lead smelting process. In this study, we adopted the MetCal v7.81 [21,22], which features a modular construction for thermodynamic models, to rapidly develop a thermodynamics simulation system with visual operation interfaces. This simulation system integrates the process flowchart and the calculation flowchart, providing an intuitive visualization of the bottom-blowing lead oxidation smelting process. It is expected to provide a useful tool for thermodynamic analysis.
In this study, we used the chemical equilibrium constant method and the MetCal v7.81 to develop a multi-phase equilibrium thermodynamic model and simulation system for the oxygen-enriched bottom-blowing lead oxidation smelting process. This process is based on the multi-phase equilibrium principle and the mechanism of the bottom-blowing lead oxidation smelting process. We simulated the equilibrium product composition and key technical indicators and compared them with industrial data in order to verify the accuracy of the model. It is expected to provide an effective tool for predicting the output of the bottom-blowing lead oxidation smelting process, revealing the impurity distribution behavior pattern, and optimizing the process parameters.

2. Modeling Principles

2.1. Lead Bottom-Blowing Oxidation Smelting Process

The lead smelting process has three stages: oxidation smelting, reduction smelting, and slag fuming. We studied the bottom-blowing lead oxidation smelting process. The oxygen-enriched bottom-blowing lead smelting process took place in a bottom-blowing furnace. The main structure is shown in Figure 1. After mixing and granulation, the carbide slag, lead glass, lead concentrate, lead mud, zinc leaching residue, return dust, and coal were added to the furnace from the air-sealed feeding port above the furnace at a certain ratio. A specific ratio of industrial oxygen was injected into the molten pool from the oxygen lance at the bottom of the furnace. The injected oxygen formed a large number of microbubbles, which entered the molten pool and dispersed throughout the melt. At the same time, the injected oxygen played a stirring role, forming good heat and mass transfer conditions. This enables the reaction to quickly approach an equilibrium state [13,14,17]. The granular material added from the feed opening fell onto the surface of the molten pool and was quickly swept into the molten pool. Under high oxygen potential, a part of lead oxide and lead sulfide interacted to form metal lead. Part of the lead oxide and silica reacted to form slag. The sulfides of iron and zinc were oxidized, and then, the metal oxides reacted with silica and calcium oxide to form the slag. The melted lead and lead-rich slag formed a lead layer and a slag layer in the molten pool because of the different densities. Finally, we obtained crude lead containing less than 0.5% S, lead-rich slag containing about 40% Pb, and flue gas containing 8–12% SO2. Part of the melt was lifted up by a high-pressure jet stream, forming oxidized dust. The furnace was kept at a specific negative pressure so that the high-temperature flue gas could enter the waste heat boiler. The crude lead was discharged into the refining furnace through the crude lead port. The slag port released the lead-rich slag into the reduction furnace.

2.2. Modeling Assumptions

According to the reaction mechanism of the bottom-blowing lead oxidation smelting process, the products of the lead concentrate oxidation smelting process consisted of primary crude lead, lead-rich slag, flue gas, and dust. In constructing a multi-phase equilibrium model for oxidation smelting, the equilibrium product consisted of only primary crude lead, lead-rich slag, and flue gas, while the dust consisted of mechanical dust and flue gas cooling dust. We assumed that the chemical composition of the product was at equilibrium, which is shown in Table 1.

2.3. Model Construction

According to the chemical equilibrium constant method, also known as the K-value method, the system reaches equilibrium at constant temperature and pressure. We established a nonlinear system of equations based on the chemical equilibrium constant equation and the mass conservation equation for each element. Then, we solved the nonlinear system of equations algorithmically to obtain the amount of each component of each phase in the system.
From the model assumptions and the phase rule:
P + F = C + 2
where P is the number of phases, F is the degree of freedom, and C is the number of components.
The bottom-blowing lead oxidation smelting system has 22 components and 60 chemical species. The system has a set of linearly independent molecular formula vectors, which are called independent species, and the remaining species are called dependent species. We selected the elements Pb, Zn, Cu, Fe, S, As, Sb, Bi, Ag, Au, Cd, Mg, Al, K, Na, Ba, O, C, H, and N and the oxides SiO2 and CaO as the ‘‘components,” which formed the stoichiometry matrix for the 22 independent species shown in Table A1 of Appendix A and the stoichiometry matrix for the 38 dependent species shown in Table A2 of Appendix A. The stoichiometric coefficient matrix of the reactions could be obtained from the stoichiometry matrix of the independent species and dependent species. The chemical reactions consisting of a linearly independent set of row vectors in the stoichiometric coefficient matrix of reactions formed the independent reaction equations. Assuming that the number of elements contained in the reaction system was Ne and the number of compound species was Nc, the amount of each species within the system was determined by independent reaction equations, where the independent reaction number Nb is equal to Nc−Ne and the Nb independent reactions can be expressed as follows:
V j , i A i , k = B j , k
where Vj,i is the stoichiometric coefficient matrix of reactions, Ai,k and Bj,k are the matrices of the constituent components of the independent species and the dependent species, respectively; and i, j, and k denote the independent species number, the dependent species number, and the species number, respectively.
According to the rules of matrix operations, Vj,i can be obtained as follows:
V j , i = B j , k U k , i
where (Uk,i) denotes the inverted matrix of (Ai,k) and must be calculable. i ∈ [1, 22], k ∈ [1, 22], and j ∈ [1, 38].
The chemical equilibrium constants Kj for the 38 independent reactions (Table 2) of the 38 dependent species produced from the 22 independent species are given by
K j = exp Δ G bj 0 V ji Δ G ai 0 RT
where R is the molar gas constant, T is the equilibrium temperature of the system,   Δ G ai 0 is the standard Gibbs free energy of formation of ith independent species, and Δ G bj 0 is the standard Gibbs free energy of formation of the jth dependent species.
When the bottom-blowing lead oxidation smelting reaction system reaches equilibrium, the resulting relationship between the 22 independent species and the 38 dependent species can be expressed as follows:
Y j = Z m , j γ j K j i γ i X i Z m , i V j , i
where X i is the mole number of the ith independent species; γ i is the activity factor of the ith independent species; Z m , i is the mole number of the phase to which the ith independent species belongs; Y j is the mole number of the jth dependent species; γ j is the activity factor of the jth dependent species; and Z m , j is the mole number of the phase to which the jth dependent species belongs.
The total mole number of the m product phase in Equation (5), Zm, is given by
Z m = i ( m ) X i + j ( m ) Y i
where i(m) means that only the independent species i that belongs to the product phase m is included in the summation; similarly, j(m) means that only the dependent species j that belongs to the product phase m is included in the summation.
According to the law of conservation of mass, the mole number of each element is calculated as follows:
Q k = i A i , k X i + j B j , k Y i
where   Q k is the mole number of element k.
The amount of each species in each phase at the equilibrium of this system can be obtained by solving the system of nonlinear equations consisting of Equations (5)–(7) according to the Newton–Raphson algorithm.

3. Basic Data and Simulation System

3.1. Raw Materials and Their Compositions

The main raw materials input included carbide slag, lead glass, lead concentrate, lead mud, zinc leaching slag, return dust, coal, air, and industrial oxygen (oxygen volume concentration of 95%). Based on the elements analysis data of raw material in the factory and the composition of common industrial raw materials [23,24,25], models for each raw material were constructed using the MetCal v7.81 to calculate their composition, as shown in Table 3, Table 4, Table 5, Table 6, Table 7, Table 8 and Table 9.

3.2. Thermodynamic Basic Data

The Gibbs free energy of equilibrium product phase components in the bottom-blowing lead oxidation smelting process can be calculated by Equation (8), which is derived from Kirchhoff’s formula and the relationship between standard molar reaction entropy and temperature in Appendix B. The standard Gibbs free energy and other relevant thermodynamic parameters for each product component were obtained through the MetCal v7.81, as shown in Table A3 of Appendix A. To eliminate the influence of reaction kinetics, the product phase activity coefficient was corrected based on the literature [11,17,18,19,26], as shown in Table A4 of Appendix A. The activity factor of the components in flue gas was 1.
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. System Development

The model described in Section 2.3 was calculated according to the algorithm flowchart shown in Figure 2. Then, according to Equation (9), we considered the thermal equilibrium of this smelting process and applied the MetCal v7.81 to develop a thermodynamic simulation system for bottom-blowing lead oxidation smelting, as shown in Figure 3.
i n A Δ H 298 , A i + i n A 298 T i c p , A i dT = j n B Δ H 298 , B j + j n B 298 T j c p , B j dT + Q Loss
where Ai is the ith reactant; nA is the amount of reactant; Ti is the initial temperature of the reactant Ai; Bj is the jth product; nB is the amount of product; Tj is the temperature of product Bj; ∆H is the enthalpy; cp is the specific heat at constant pressure; and QLoss is the heat loss.

4. Model Validation

4.1. Calculation Conditions

We used the developed thermodynamic simulation system for bottom-blowing lead oxidation smelting to calculate the equilibrium product composition for a lead smelter in China. We used their average operating parameters from December 2021 to February 2022 as the calculating conditions. The total feedstock input was 105 t/h, which was granulated by 65.5% lead concentrate, 2% calcium carbide slag, 3.5% lead mud, 6% zinc leach slag, 20% return dust, 0.5% lead glass, and 2.5% coal. The oxygen-enriched concentration was 90%. The oxygen/feed ratio was 110 Nm3/t, and the melting temperature was obtained from heat balance calculations. We assumed that the primary crude lead temperature was 50 °C lower than the lead-rich slag temperature and the flue gas and dust temperature was 50 °C higher than the lead-rich slag temperature.

4.2. Calculation Result Verification

Based on the industrial data of a Chinese lead smelter from December 2021 to February 2022, the industrial elements analysis data of primary crude lead, lead-rich slag, and dust were averaged and compared with the model calculated data. The results are shown in Figure 4.
The calculated values of the elements in the product, excluding some elements that were not detected in production, closely matched the industrial data. The relative errors of the elemental compositions of Pb, Zn, Cu, S, As, Sb, Bi, and Cd in the primary crude lead were 0.18%, 6.25%, 0.66%, 5.92%, 3.38%, 7.65%, 0.97%, and 3.33%, respectively. The relative errors of the elemental compositions of Pb, Zn, Cu, S, FeO, SiO2, CaO, Mg, Al, As, and Sb in the lead-rich slag were 1.95%, 5.23%, 4.83%, 6.05%, 2.03%, 3.09%, 5.80%, 7.14%, 5.95%, 1.12%, and 0.90%, respectively. The relative errors of the compositions of Pb, S, ZnO, As, Sb, Ag and Cd in dust were 2.83%, 3.94%, 2.60%, 1.48%, 2.78%, 4.17% and 3.45%, respectively.
According to the results of the data presented in Figure 4, the relative error of element Sb in primary crude lead was the largest, 7.65%. In lead-rich slag, the relative error of Mg was the largest, 7.14%, and the relative error of Ag in dust was the largest, 4.17%. The relative errors of some elements are relatively large, which may be caused by three factors. Firstly, there is a lack of thermodynamic parameters for some elements in the high-temperature smelting system of lead, and the deviation of activity coefficients affects the accuracy of the calculated results. Secondly, the model assumed that the oxidation smelting process is at a constant temperature, but in reality, the temperature of the oxidation smelting process is in a dynamic state of change. Finally, during the iterative calculation of the model, errors will gradually accumulate, and these errors will gradually accumulate as iterations increase. The average relative error of all elements was 3.76%. The error range of these results was small, and the results showed that the model reflected the smelting characteristics of bottom-blowing lead oxidation and provided a useful tool for subsequent thermodynamic analysis of the system.
The calculated values of key technical indicators, such as the iron–silica ratio and the calcium–silica ratio in slag, soot rate, primary crude lead productivity, and lead-rich slag temperature, were 1.98, 0.46, 15.4%, 21.01%, and 1030 °C, respectively, and the corresponding production averages were 1.90, 0.45, 15%, 24%, and 1150 °C, with relative errors of 4.21%, 2.22%, 2.66%, 12.45%, and 10.43%, respectively. The main reasons for the deviations in the primary crude lead productivity and the temperature of the lead-rich slag are as follows: the heat transfer behavior during the lead oxidation smelting process is very complex, including convection, diffusion, radiation, and other heat transfer modes between the furnace shell and furnace lining, as well as between the outer wall of the furnace shell and the outer environment. Therefore, there were errors in the calculation of thermal equilibrium, which in turn led to errors in the primary crude lead productivity. The results obtained by the constructed thermodynamic simulation system for the bottom-blowing lead oxidation smelting process closely reflected the actual production situation and reflected the bottom-blowing lead oxidation smelting production process, which can be used as an effective tool for subsequent industrial production parameter optimization.

5. Conclusions

(1)
Based on the mechanism of the bottom-blowing lead oxidation smelting process, we constructed a multi-phase equilibrium mathematical model of the bottom-blowing lead oxidation smelting process using the chemical equilibrium constant method. We developed a bottom-blowing lead oxidation smelting thermodynamic simulation system based on the MetCal v7.81 and provided an effective tool for the thermodynamic calculation of the process.
(2)
We validated the model using the average operating parameters of the bottom-blowing lead oxidation smelting of a Chinese lead smelter as the calculation conditions. The average relative error of the calculation value of the mass fraction in the crude lead, lead-rich slag, and dust was 3.76%. The results closely matched the production values, indicating that the model could reflect the multi-phase reaction characteristics of the bottom-blowing lead oxidation smelting process. Therefore, it is a useful tool for the subsequent thermodynamic analysis of the system.
(3)
The calculated values of the key technical indicators of the lead bottom-blowing smelting process had small errors with the average measured values of industrial production. The average relative error of important technical indicators, including dust rate, crude lead productivity, lead-rich slag temperature, slag iron–silica ratio (RFe/SiO2), and slag calcium–silica ratio (RCaO/SiO2), was 6.39%, which indicated that the constructed model closely reflected the actual lead bottom-blowing smelting process and had the potential to improve the production process and optimize the process parameters.

Author Contributions

Conceived and designed the experiments, M.L. and M.Y.; performed the experiments, M.Y. and X.C.; analyzed the data, M.Y. and X.C.; searched the relevant literature and data, M.Y., X.C. and M.L.; wrote the paper, X.C.; reviewed and contributed to the final manuscript, F.L., J.H., M.L. and X.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program (No. 2022YFC2904201), the Natural Science Foundation of Jiangxi Province of China (No. 20212BAB204026), the China Baowu Low Carbon Metallurgy Innovation Foundation (No. BWLCF202121), the China Postdoctoral Science Foundation (No. 2019M662268), National Nature Science Foundation of China (No. 52264047), Training Plan for Academic and Technical Leaders of Major Disciplines in Jiangxi Province (20225BCJ23009), Natural Science Foundation for Distinguished Young Scholars of Jiangxi Province (No.20232ACB214006), the Postdoctoral program of Jiangxi Province (No. 2018KY15), the Jiangxi Province Postdoctoral Daily Funding Project (No. 2019RC28).

Data Availability Statement

Data available on request from the authors.

Acknowledgments

The author is grateful to Qing Shu and Yanxin Wu for their advice on this paper. Special thanks to Henan Yuguang Gold Lead Co., Ltd. for providing the production data for this research.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Stoichiometry matrix for 22 independent species.
Table A1. Stoichiometry matrix for 22 independent species.
Com.PhasePbZnCuFeSAsSbBiAgAuCdMgAlKNaBaSiO2CaOOCHN
PbLd1000000000000000000000
PbSLd1000100000000000000000
ZnLd0100000000000000000000
CuLd0010000000000000000000
FeLd0001000000000000000000
AsLd0000010000000000000000
SbLd0000001000000000000000
BiLd0000000100000000000000
AgLd0000000010000000000000
AuLd0000000001000000000000
CdLd0000000000100000000000
PbOSl1000000000000000001000
PbSiO3Sl1000000000000000101000
CaOSl0000000000000000010000
MgOSl0000000000010000001000
AlO1.5Sl0000000000001000001.5000
KO0.5Sl0000000000000100000.5000
NaO0.5Sl0000000000000010000.5000
BaSO4Sl0000100000000001004000
COGas0000000000000000001100
N2Gas0000000000000000000002
H2OGas0000000000000000001020
Table A2. Stoichiometry matrix for 38 dependent species.
Table A2. Stoichiometry matrix for 38 dependent species.
Com.PhasePbZnCuFeSAsSbBiAgAuCdMgAlKNaBaSiO2CaOOCHN
ZnSLd0100100000000000000000
CuS0.5Ld01000.500000000000000000
FeSLd0001100000000000000000
PbSSl1000100000000000000000
PbSO4Sl1000100000000000004000
ZnOSl0100000000000000001000
Zn2SiO4Sl0200000000000000102000
ZnFe2O4Sl0102000000000000004000
CuO0.5Sl0010000000000000000.5000
CuS0.5Sl00100.500000000000000000
CuFe2O4Sl0012000000000000004000
FeOSl0001000000000000001000
FeO1.33Sl0001000000000000001.33000
FeSSl0001100000000000000000
2FeO·SiO2Sl0002000000000000102000
SiO2Sl0000000000000000100000
AsO1.5Sl0000010000000000001.5000
SbO1.5Sl0000001000000000001.5000
BiO1.5Sl0000000100000000001.5000
CdOSl0000000000100000001000
AgO0.5Sl0000000010000000000.5000
AuSl0000000001000000000000
O2Gas0000000000000000002000
PbGas1000000000000000000000
PbOGas1000000000000000001000
PbSGas1000100000000000000000
ZnGas0100000000000000000000
ZnOGas0100000000000000001000
ZnSGas0100100000000000000000
CdOGas0000000000100000001000
CdSGas0000100000100000000000
SO2Gas0000100000000000002000
S2Gas0000200000000000000000
CO2Gas0000000000000000002100
AsO1.5Gas0000010000000000001.5000
AsS1.5Gas00001.510000000000000000
SbO1.5Gas0000001000000000001.5000
SbS1.5Gas00001.501000000000000000
Table A3. Thermodynamic parameters of species.
Table A3. Thermodynamic parameters of species.
ComponentState Δ H 298 θ /
(kJ·mol−1)
Δ S 298 θ /
(J·K−1·mol−1)
cp = a + b × 10−3T + c × 105T−2 + d × 10−6T2
abcd
PbLiquid3.87370.50627.1591.02900
ZnLiquid5.72748.54931.381000
CuLiquid8.02834.23632.845000
FeLiquid8.00623.52140.8791.67400
SLiquid032.07132.005−0.002−0.0380
AsLiquid21.56853.28428.833000
SbLiquid17.53162.71231.381000
BiLiquid9.27171.98027.197000
CdLiquid5.60760.71729.707000
AuLiquid047.489−268.634237.1391418.47−52.813
AgLiquid6.39343.22033.473000
PbOLiquid−202.24973.37965.000000
PbSLiquid−93.14384.12966.946000
ZnOLiquid−309.54247.92060.669000
ZnSLiquid−203.00558.66149.7534.448−4.551−0.005
AsO1.5Liquid−643.439128.125152.720000
SbO1.5Liquid−675.490143.628156.904000
BiO1.5Liquid−578.024149.814202.005000
CdOLiquid−258.99654.81247.2646.364−4.9080
PbGas195.205175.37728.063−11.029−9.3104.728
PbOGas68.139240.04841.612−3.526−20.1361.014
PbSGas127.959251.41637.3500.194−2.0960.140
ZnGas130.40316.99220.898−0.133−0.0670.034
ZnOGas136.518242.81137.671−0.286−1.9850.735
ZnSGas204.322236.404166.350−85.742−166.12521.952
AsO1.5Gas−322.845371.92582.1346.444−5.3560
AsS1.5Gas27.042314.28996.2011.071−8.2130
SbO1.5Gas−708.564129.903180.004000
SbS1.5Gas119.661409.820107.6360.209−7.2550
CdOGas127.003231.57043.560−10.649−11.8195.291
CdSGas175.662244.98736.257−3.86718.6113.678
O2Gas0205.15434.8601.312−14.1410.163
SO2Gas−296.820248.22654.7813.350−24.745−0.241
S2Gas128.603228.16934.6723.286−2.816−0.312
COGas−110.544197.66529.9325.415−10.814−1.054
CO2Gas−393.515213.77454.4375.116−43.579−0.806
N2Gas0191.61523.52912.1171.210−3.076
H2OGas−241.832188.83731.43814.106−24.952−1.832
Table A4. Activity factor of species.
Table A4. Activity factor of species.
Com.PhaseActivity FactorCom.PhaseActivity FactorCom.PhaseActivity Factor
PbLd0.0196PbOSl0.0036PbOSl0.0036
PbSLd80PbSSl50PbSSl50
ZnLd150PbSO4Sl0.01PbSO4Sl0.01
ZnSLd1450PbSiO3Sl0.01PbSiO3Sl0.01
CuLd0.3ZnOSl1ZnOSl1
CuS0.5Ld40Zn2SiO4Sl1Zn2SiO4Sl1
FeLd10ZnFe2O4Sl1ZnFe2O4Sl1
FeSLd10CuO0.5Sl1CuO0.5Sl1
AsLd50CuS0.5Sl1AsO1.5Sl1
BiLd4CuFe2O4Sl1BiO1.5Sl1
SbLd7FeOSl1SbO1.5Sl1
AgLd1FeO1.33Sl0.1NaO0.5Sl1
AuLd1FeSSl1AgO0.5Sl1
CdLd82FeO·SiO2Sl1AuSl1
MgOSl1CaOSl1CdOSl1
AlO1.5Sl1BaSO4Sl1OtherSl1
SiO2Sl1KO0.5Sl1

Appendix B

The derivation of Equation (8) is as follows:
Δ H T θ = Δ H 298 θ + 298 T c P dT
Δ S T θ = Δ S 298 θ + 298 T c P T dT
Δ G T θ = Δ H T θ T · Δ S 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

References

  1. Jiang, J.M. Situation and Development Trend of Lead Smelting Technology at Home and Abroad. Energy Sav. Nonferrous Met. 2013, 29, 4–8. (In Chinese) [Google Scholar]
  2. Chen, L.; Hao, Z.D.; Yang, T.Z.; Liu, W.F.; Zhang, D.C.; Zhang, L.; Bin, S.; Bin, W.D. A Comparison Study of the Oxygen-Rich Side Blow Furnace and the Oxygen-Rich Bottom Blow Furnace for Liquid High Lead Slag Reduction. JOM 2015, 67, 1123–1129. [Google Scholar] [CrossRef]
  3. Queneau, P.E. The QSL reactor for lead and its prospects for Ni, Cu and Fe. JOM 1989, 41, 30–35. [Google Scholar] [CrossRef]
  4. Chen, L.; Yang, T.Z.; Bin, S.; Liu, W.F.; Zhang, D.C.; Bin, W.D.; Zhang, L. An Efficient Reactor for High-Lead Slag Reduction Process: Oxygen-Rich Side Blow Furnace. JOM 2014, 66, 1664–1669. [Google Scholar] [CrossRef]
  5. Zhukov, V.P.; Kolmachikhin, B.V.; Kolmachikhina, É.B. Hydrogasdynamic Regularities of Melt Blowing in “Ausmelt” Furnace by Cold Modeling Methods. Metallurgist 2020, 63, 1220–1226. [Google Scholar] [CrossRef]
  6. Ramus, K.; Hawkins, P. Lead/acid battery recycling and the new Isasmelt process. J. Power Sources 1993, 42, 299–313. [Google Scholar] [CrossRef]
  7. Errington, W.J.; Edwards, J.S.; Hawkins, P. Isasmelt technology-current and future development. J. South Afr. Inst. Min. Met. 1997, 97, 161–167. [Google Scholar]
  8. Li, J.D.; Song, Y.P.; Zhou, P.; Ma, J.; Chen, Z.; Chai, L.Y. A RQPSO Algorithm for Multiphase Equilibrium Calculation in the KIVCET Process. JOM 2018, 70, 2893–2899. [Google Scholar] [CrossRef]
  9. Gu, Y.; Wang, J.P.; Wang, X.; Che, D.; Wu, Q.X. Situation and sustainable development proposals for lead resource development in China. Resour. Ind. 2018, 20, 39–46. (In Chinese) [Google Scholar] [CrossRef]
  10. Wang, Q.M.; Huang, M.X.; Wang, S.S.; Li, D.; Tian, Q.H.; Guo, X.Y. Research Progress on Thermodynamic Model of Typical Nonferrous Metal Smelting Process. Nonferrous Met. 2021, 10, 1–12. (In Chinese) [Google Scholar]
  11. Nagamori, M.; Chaubal, P.C. Thermodynamics of copper matte converting: Part III. Steady-state volatilization of Au, Ag, Pb, Zn, Ni, Se, Te, Bi, Sb, and As from slag, matte, and metallic copper. Met. Trans. B 1982, 13, 319–329. [Google Scholar] [CrossRef]
  12. Wang, Q.M.; Guo, X.Y.; Tian, Q.H.; Jiang, T.; Chen, M.; Zhao, B.J. Development and Application of SKSSIM Simulation Software for the Oxygen Bottom Blown Copper Smelting Process. Metals 2017, 7, 431. [Google Scholar] [CrossRef]
  13. Liu, F.K. Numerical Simulation and Optimization of Multiphase Flow in Bottom Blowing Lead Smelting Furnace. Master’s Thesis, Central South University, Changsha, China, 2013. [Google Scholar]
  14. Song, K.Z.; Jokilaakso, A. CFD Modeling of Multiphase Flow in an SKS Furnace with New Tuyere Arrangements. Met. Mater. Trans. B Proc. Met. Mater. Proc. Sci. 2022, 53, 253–272. [Google Scholar] [CrossRef]
  15. Shimpo, R.; Goto, S.; Ogawa, O.; Asakura, I. A Study on the Equilibrium between Copper Matte and Slag. Can. Met. Q. 1986, 25, 113–121. [Google Scholar] [CrossRef]
  16. Li, M.Z.; Zhou, J.M.; Tong, C.R.; Zhang, W.H.; Chen, Z.; Wang, J.L. Thermodynamic Modeling and Optimization of the Copper Flash Converting Process Using the Equilibrium Constant Method. Met. Mater. Trans. 2018, 49, 1794–1807. [Google Scholar] [CrossRef]
  17. Chen, L.; Wang, Z.H.; Chen, W.; Xiao, H.; Liu, W.F.; Zhang, D.C.; Yang, T.Z. Thermodynamic Simulation on Elements Distribution of Lead Concentrate Oxidative Smelting in Oxygen-Rich Bottom-Blow Smelting Process. Nonferrous Met. Extr. Met. 2018, 9, 1–6. (In Chinese) [Google Scholar] [CrossRef]
  18. Tan, P.F.; Zhang, C.F.; Zhang, R.Y. A Computer Model of QSL Lead Smelting Process. J. Cent. South Univ. Sci. Technol. 1996, 5, 40–43. (In Chinese) [Google Scholar]
  19. Zhang, C.F.; Tan, P.F.; Zeng, D.W.; Zhao, T.L. Computer simulation and theoretical analysis of QSL lead smelting process. Nonferrous Met. Extr. Met. 1997, 1, 13–16. (In Chinese) [Google Scholar]
  20. Shishin, D.; Hidayat, T.; Sultana, U.; Shevchenko, M.; Jak, E. Experimental measurement and thermodynamic model predictions of the distributions of Cu, As, Sb and Sn between liquid lead and PbO–FeO–Fe2O3–SiO2 slag. Int. J. Mater. Res. 2020, 111, 733–743. [Google Scholar] [CrossRef]
  21. Chun, W.X.; Liang, W.J.; Qing, W.H. Research on the Mathematical Model of Local Equilibrium in the Top-Blown Smelting Process of Electronic Waste. Metals 2021, 11, 1500. [Google Scholar] [CrossRef]
  22. Li, M.Z.; Zhou, J.M.; Tong, C.R.; Zhang, W.H.; Li, H.S. Mathematical model of whole-process calculation for bottom-blowing copper smelting. Met. Res. Technol. 2018, 115, 107. [Google Scholar] [CrossRef]
  23. Song, J.B.; Xi, W.L.; Zhu, L.P.; Liu, S.H. Experimental study on oxygen-blowing co-smelting and desulfurization of lead concentrate with lead paste and lead glass. China Nonferrous Met. 2022, 51, 110–117. (In Chinese) [Google Scholar] [CrossRef]
  24. Li, D.B. Morder Oxygen Bottom Blowing Lead Smelting Technology; Metallurgical Industry Press: Beijing, China, 2020; pp. 71–76. (In Chinese) [Google Scholar]
  25. Yao, G.; Guo, Q.; Li, Y.; Xu, Z.; Han, Z.; He, M.; Qi, T. An innovation technology for recovering silver and valuable metals from hazardous zinc leaching residue through direct reduction. Miner. Eng. 2022, 188, 107857. [Google Scholar] [CrossRef]
  26. 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. (In Chinese) [Google Scholar]
Figure 1. Schematic diagram of the structure of a bottom-blowing lead oxidation furnace.
Figure 1. Schematic diagram of the structure of a bottom-blowing lead oxidation furnace.
Processes 11 03043 g001
Figure 2. Calculation flowchart of thermodynamic model.
Figure 2. Calculation flowchart of thermodynamic model.
Processes 11 03043 g002
Figure 3. Thermodynamic simulation system for bottom-blowing lead oxidation smelting.
Figure 3. Thermodynamic simulation system for bottom-blowing lead oxidation smelting.
Processes 11 03043 g003
Figure 4. Relative error of product element composition between calculated values and industrial data.
Figure 4. Relative error of product element composition between calculated values and industrial data.
Processes 11 03043 g004
Table 1. The chemical composition of the product.
Table 1. The chemical composition of the product.
PhasesChemical Components
Primary crude lead (Ld)Pb, PbS, Zn, ZnS, Cu, CuS0.5, Fe, FeS, As, Bi, Sb, Ag, Au, Cd, Others;
Lead-rich slag (Sl)PbO, PbS, PbSO4, PbSiO3, ZnO, Zn2SiO4, ZnFe2O4, CuO0.5, CuS0.5, CuFe2O4, FeO, FeO1.33, FeS, 2FeO·SiO2, CaO, MgO, AlO1.5, SiO2, KO0.5, NaO0.5, BaSO4, AsO1.5, BiO1.5, SbO1.5, CdO, AgO0.5, Au, Others;
Flue gas (Gas)O2, Pb, PbO, PbS, Zn, ZnO, ZnS, CdO, CdS, SO2, S2, CO, CO2, N2, AsO1.5, AsS1.5, SbO1.5, SbS1.5, H2O;
Dust (Dt)Ca(OH)2, CaCO3, CaSO4, SiO2, AlO1.5, FeO1.33, MgSiO3, K2SiO3, Na2SiO3, H2O, PbO·SiO2, CaSiO3, PbS, PbSO4, ZnS, CuS0.5, FeS2, FeO1.5, AsS1,5, SbS1.5, BiS1.5, CdS, AgO0.5, Au, PbO, BaSO4, ZnSO4, FeSO4, ZnO, CuO, FeO, CaO, Ag, AsO1.5, SbO1.5, BiO1.5, CdO, PbSiO3, Zn2SiO4, Ag2SO4, CuO0.5, Pb, Zn, Cu, Fe, FeS, As, Bi, Sb, Cd, ZnFe2O4, CuFe2O4, 2FeO•SiO2, MgO, KO0.5, NaO0.5, Others;
Table 2. Equilibrium reactions for the bottom-blowing lead oxidation smelting process.
Table 2. Equilibrium reactions for the bottom-blowing lead oxidation smelting process.
Equilibrium ReactionKjEquilibrium ReactionKj
Pb(Ld) + 0.5O2(gas) = PbO(Sl)K1Zn(gas) + 0.5O2(gas) = ZnO(gas)K20
PbS(Ld) + O2(gas) = Pb(Sl) + SO2(gas)K2ZnS(gas) + 1.5O2(gas) = ZnO(gas) + SO2(gas)K21
Zn(Ld) + 0.5O2(gas) = ZnO(Sl)K3S2(gas) + 2O2(gas) = 2SO2(gas)K22
CuS0.5(Sl) + 1.5O2(gas) = CuO0.5(Sl) + 0.5SO2(gas)K4CO(gas) + 0.5O2(gas) = CO2(gas)K23
FeS(Ld) + 1.5O2(gas) = FeO(Sl) + SO2(gas)K5FeO(Sl) + 0.165O2(gas) = FeO1.33(Sl)K24
Pb(Ld) = Pb(gas)K6FeS(Sl) + 1.5O2(gas) = FeO(Sl) + SO2(gas)K25
Cd(Ld) + 0.5O2(gas) = CdO(Sl)K7CuS0.5(Sl) + 0.75O2(gas) = CuO0.5(Sl) + 0.5SO2(gas)K26
Pb(gas) + 0.5O2(gas) = PbO(gas)K8PbS(Sl) = PbS(gas)K27
PbS(Sl) + 2PbO(Sl) = 3Pb(Ld) + SO2(gas)K9CdO(Sl) = CdO(gas)K28
AsO1.5(Sl) = AsO1.5(gas)K10FeO(Sl) + CO(gas) = Fe(Ld) + CO2(gas)K29
AsS1.5(gas) + 2.25O2(gas) = AsO1.5(gas) + 1.5SO2(gas)K11ZnS(Ld) + 1.5O2(gas) = ZnO(Sl) + SO2(gas)K30
As(Ld) + 0.75O2(gas) = AsO1.5(Sl)K12ZnO(Sl) + 0.5SiO2(Sl) = 0.5Zn2SiO4(Sl)K31
Bi(Ld) + 0.75O2(gas) = BiO1.5(Sl)K13Zn(Ld) = Zn(gas)K32
SbO1.5(Sl) = SbO1.5(gas)K14CuO0.5(Sl) + 0.5CO(gas) = Cu(Ld) + 0.5CO2(gas)K33
SbS1.5(gas) + 2.25O2(gas) = SbO1.5(gas) + 1.5SO2(gas)K15Au(Ld) = Au(Sl)K34
Sb(Ld) + 0.75O2(gas) = SbO1.5(Sl)K16PbO(Sl) + SiO2(Sl) = PbSiO3(Sl)K35
AgO0.5(Ld) + 0.5CO(gas) = Ag(Ld) + 0.5CO2(gas)K17FeO(Sl) + SiO2(Sl) = FeO·SiO2(Sl)K36
PbS(Sl) + PbSO4(Sl) = 2Pb(Ld) + 2SO2(gas)K18ZnO(Sl) + 2FeO1.33(Sl) + 0.167O2(gas) = ZnFe2O4(Sl)K37
CdS(gas) + 1.5O2(gas) = CdO(gas) + SO2(gas)K19CuO0.5(Sl) + 2FeO(Sl) + 0.75O2(gas) = CuFe2O4(Sl)K38
Table 3. Chemical composition of carbide slag (wt%).
Table 3. Chemical composition of carbide slag (wt%).
Ca(OH)2CaCO3CaSO4SiO2Al2O3Fe3O4
51.560.0220.251.140.510.20
MgSiO3K2SiO3Na2SiO3H2OOther
1.490.0890.183312.87
Table 4. Chemical composition of lead concentrate (wt%).
Table 4. Chemical composition of lead concentrate (wt%).
PbSPbSO4ZnSCu2SFeS2Fe2O3CaCO3SiO2
51.16.5.197.881.4915.592.572.235.07
As2S3Sb2S3Bi2S3CdSAg2OAuH2OOther
0.730.400.250.030.150.00037.20.0597
Table 5. Chemical composition of zinc leaching slag (wt%).
Table 5. Chemical composition of zinc leaching slag (wt%).
PbSPbSO4PbSiO3ZnSO4Zn2SiO4ZnOZnSCdO
7.2432.124.286.235.712.432.921.11
Fe3O4CaSO4SiO2Al2O3AgSO4H2OOther
0.393.300.221.070.04230.94
Table 6. Chemical composition of return dust (wt%).
Table 6. Chemical composition of return dust (wt%).
PbSPbOPbSO4ZnSO4ZnOCuOFeO
0.335.3450.8532.402.330.030.37
CaOSiO2AgAuAs2O3As2S3Sb2O3
0.30.1000.090.550.31
Sb2S3Bi2O3Bi2S3CdOCdSOther
0.050.390.065.890.440.17
Table 7. Chemical composition of lead mud (wt%).
Table 7. Chemical composition of lead mud (wt%).
PbOPbSO4BaSO4CaSO4ZnSO4FeSO4SiO2H2OOther
18.3854.041.520.520.220.040.271015.01
Table 8. Chemical composition of lead glass (wt%).
Table 8. Chemical composition of lead glass (wt%).
PbS·SiO2SiO2CaSiO3K2SiO3Na2SiO3MgSiO3H2OOther
33.1819.889.3616.268.424.4308.47
Table 9. Chemical composition of coal (wt%).
Table 9. Chemical composition of coal (wt%).
CCH4CO2H2N2H2SFe2O3
805.051.880.981.150.531.00
CaOMgOAl2O3H2OOtherSiO2
0.240.053.280.390.644.81
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

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. https://0-doi-org.brum.beds.ac.uk/10.3390/pr11103043

AMA Style

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(10):3043. https://0-doi-org.brum.beds.ac.uk/10.3390/pr11103043

Chicago/Turabian Style

Chen, Xinzhou, Mingzhou Li, Fupeng Liu, Jindi Huang, and Minghao Yang. 2023. "Multi-Phase Equilibrium Model of Oxygen-Enriched Lead Oxidation Smelting Process Based on Chemical Equilibrium Constant Method" Processes 11, no. 10: 3043. https://0-doi-org.brum.beds.ac.uk/10.3390/pr11103043

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