Next Article in Journal
Effect of Laser Welding Parameters with Different Fillers on Solidification Cracking and Mechanical Properties of AA7075
Previous Article in Journal
Probing Localised Corrosion Inhibition of AA2024-T3 by Integrating Electrode Array, SVET, SECM, and SEM-EDS Techniques
Previous Article in Special Issue
Effect of Al2O3 Inclusions or Mold Flux Particles on Their Surrounding Microstructures of Sliver Defects on the Surface of Automobile Exposed Panel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Activation Energy of Alumina Dissolution in FeO-Bearing Slags

1
Department of Advanced Materials Engineering, Technical University of Korea, 237, Sangidaehak-ro, Shiheung-si 15073, Republic of Korea
2
R&D Center, Dongkuk Steel Company, 70, Geonposaneop-ro 3214beon-gil, Nam-gu, Pohang-si 37874, Republic of Korea
*
Author to whom correspondence should be addressed.
Metals 2023, 13(10), 1702; https://doi.org/10.3390/met13101702
Submission received: 31 August 2023 / Revised: 26 September 2023 / Accepted: 27 September 2023 / Published: 6 October 2023
(This article belongs to the Special Issue Non-metallic Inclusions in Steelmaking)

Abstract

:
The dissolution of Al2O3 non-metallic inclusions in slag containing FeO was investigated in this study. The slag system used in the experiments was a quaternary system of CaO-SiO2-Al2O3-FeO. The composition of the slag was studied by fixing the basicity (CaO/SiO2 ratio) to 1 and varying the FeO content to 10 and 20 wt%. In addition, the experimental temperature was varied to 1550 °C, 1575 °C, and 1600 °C to study the effect of temperature on the Al2O3 dissolution behavior. The experimental equipment used was a single hot thermocouple apparatus. The dissolution rate of Al2O3 particles increased linearly with increasing temperature and FeO content. In addition, the mass transfer activation energy of Al2O3 dissolution in FeO 10 wt% and FeO 20 wt% was calculated through an Arrhenius-type analysis. The obtained mass transfer activation energies were 159 and 189 kJ/mole, respectively.

1. Introduction

In line with South Korea’s 2030 carbon neutrality goal, the steel industry is increasingly interested in electric arc furnaces, which emit less carbon dioxide than blast furnaces. Accordingly, research on electric arc furnace processes is necessary [1]. The electric arc furnace process injects oxygen into the steel to reduce operating time, oxidation refining, etc. [2]. The presence of oxygen in the steel can cause problems such as corrosion and hot shortness. Therefore, a deoxidation process is essential [3]. The deoxidation process is mainly carried out using aluminum, which is a strong deoxidizer [4]. When tapping molten steel, aluminum is added to the ladle furnace to deoxidize it [5]. This process produces Al2O3 inclusions, which cause several problems including fatigue failure of the steel and nozzle clogging [6]. It is therefore important to remove Al2O3 inclusions as slag.
There are two ways to remove Al2O3 inclusions. The first is Ca treatment, where the Al2O3 inclusions are removed by adding Ca to transform the Al2O3-inclusion solid phase to a liquid CaO-Al2O3 phase [7]. However, this method has the disadvantage of interaction with the bottom lining refractory, which causes corrosion of the refractory. In addition, CaS inclusions, which are as harmful as Al2O3 inclusions, are easily formed [6]. The second is to float and separate Al2O3 inclusions to the slag/metal interface in a ladle refining process and then dissolve and remove them from the top layer of slag [8]. This method has been used not only in electric furnace processes, but also in blast furnace processes. For this reason, the dissolution behavior of non-metallic inclusions in slag has been extensively studied during the past 30 years.
Sridhar et al. [9] studied the dissolution behavior of Al2O3 particles in CaO-SiO2-Al2O3-MgO slag with temperature as a variable. They reported that the dissolution of Al2O3 is dominated by mass transfer through the boundary layer. Q. Shu et al. [10] investigated the effect of Na2O addition on the dissolution behavior of cylindrical Al2O3 in CaO-Al2O3-MgO-SiO2 slag. They reported that the dissolution mechanism of Al2O3 is boundary layer mass transfer, and the dissolution rate of cylindrical Al2O3 increases with a decrease in slag viscosity and increase in thermodynamic driving force with Na2O addition. Yi. K. Wi et al. [11] investigated the dissolution behavior of Al2O3 and MgO particles within Al2O3-CaO-MgO slag, with temperature as the variable. Their research revealed that the dissolution kinetics of Al2O3 particles are primarily influenced by diffusion processes, whereas the dissolution of MgO particles adheres to chemical reaction kinetics. C. Ren et al. [12] examined how Al2O3 particles dissolve within CaO-Al2O3-SiO2 slag, considering variations in slag composition and temperature. They reported that the rate-limiting step of Al2O3 particle dissolution in CaO-Al2O3-SiO2 slag is diffusion in liquid slag, and an increase in C/A(CaO/Al2O3 ratio) and C/S(CaO/SiO2 ratio) increased the dissolution rate of Al2O3 particles. L. Holappa et al. [13] studied the dissolution behavior of Al2O3 particles and MgO⋅Al2O3 particles in CaO-SiO2-Al2O3-MgO slag with basicity as a variable. They reported that Al2O3 particles and MgO⋅Al2O3 particles dissolve slowly in slag with low basicity, and the dissolution rate of the particles increases in slag with low viscosity. H. Um et al. [5] investigated the dissolution behavior of Al2O3 particles in CaO-SiO2-Al2O3-FexO slag with FexO content as a variable. They reported that as the FexO content increased from 0 to 20 wt%, the dissolution rate of Al2O3 particles increased due to the decrease in viscosity. However, when the FexO content increased to 30 wt%, the dissolution rate did not increase due to the formation of the CA6 phase at the Al2O3 particle interface. Y. Park et al. [14] studied the dissolution behavior of wall-type Al2O3 in CaO-Al2O3-FexO-MgO-SiO2 slag with C/A and FexO content as variables. They reported that increasing temperature, increasing C/A, and increasing FexO content increased the dissolution rate of Al2O3. They also conducted temperature-variable experiments on slag with one composition to derive the mass transport activation energy of Al2O3 dissolution, which they reported to be 193.6 kJ/mole. S. Yeo et al. [15] conducted a study on the dissolution behavior of Al2O3 particles in CaO-Al2O3-SiO2 slag with Al2O3 composition and the temperature of the slag as variables. The investigators found that the dissolution rate increased with the increasing activity of Al2O3. In addition, they reported that the diffusion activation energy of Al2O3 dissolution was in a range of about 320 to 490 kJ/mole, depending on the composition of Al2O3 in the slag.
In addition, the dissolution behavior of Al2O3 in slags under various conditions has been studied [16,17,18,19,20]. However, since electric arc furnace slags use Fe scrap as raw material, the content of FeO in the slag increases [2]. For this reason, the composition of the slag used in electric furnace research must include FeO. However, there has been little research on the dissolution behavior of non-metallic inclusions in slag containing FeO. In addition, the temperature of electric arc furnace slag varies from 1550 to 1700 °C, but few studies have been conducted with temperature as a variable in slag containing FeO [21]. For this reason, the dissolution behavior of Al2O3 particles in CaO-SiO2-Al2O3-FeO slag was studied in this work with temperature and FeO content as variables.

2. Materials and Methods

2.1. Sample Preparation

Table 1 shows the chemical composition and diameter of the Al2O3 particles used in the experiment. The shape of Al2O3 particles is spherical, with a diameter of 500 ± 2.5 μm (Goodfellow Cambridge limited, Cambridgeshire, UK). The average weight of the particles was 0.25 mg, and the purity of the Al2O3 was 99.9%.
Table 2 shows the chemical composition and basicity of the slag used in the experiments. The basicity was fixed at 1, and the FeO content was varied from 10 to 20 wt%. The amount of slag used in each experiment was set to 4 mg for stability in the experiment. The slag was prepared by mixing CaO powder, prepared by calcining CaCO3 at 1200 °C with Al2O3 and FeO powder individually, and melting them in a high-frequency induction furnace.

2.2. Single Hot Thermocouple Apparatus (SHT Apparatus)

In this study, we observed the dissolution behavior of Al2O3 particles using a single hot thermocouple (SHT) apparatus. Figure 1 is a schematic diagram of the SHT apparatus. It consisted of a B-type thermocouple to melt slag and dissolve Al2O3 particles, an SHT controller to control and check the temperature of the B-type thermocouple, a video camera to observe and record the dissolution behavior of Al2O3 particles in real time, and an optical microscope.
The SHT apparatus has several advantages, including the ability to inject Al2O3 particles at the desired temperature range, quenching at 300 °C/s using the SHT controller, ease of preparing quenching specimens, and real-time observation of the melting behavior of Al2O3 particles with a video camera and optical microscope.

2.3. Experimental Conditions

Figure 2 shows the process of the Al2O3 particle dissolution experiment using the SHT apparatus. First, the + and − poles of the B-type thermocouple were welded to form an oval shape, and then mounted on the copper tip. The slag was then placed on the B-type thermocouple and the temperature was raised to 8 °C/s using an SHT controller to melt the slag. When the experimental temperature (1550, 1575, 1600 °C) was reached, Al2O3 particles were added to the melted slag to dissolve it. Subsequently, after the set time (120, 240, 360 sec), quenching was performed using the SHT controller.
As mentioned earlier, the SHT apparatus has the advantage of allowing observation of the dissolution behavior of Al2O3 particles in real time. Several studies using this apparatus have taken advantage of this by observing the dissolution behavior of inclusions in real time [5,15,22,23]. However, in this study, there was a problem, as the dissolution behavior of Al2O3 particles could not be observed in real time due to the opacity of the slag at high temperature because it contained FeO, a transition metal. To solve this problem, quenching specimens were prepared by exploiting one of the advantages of the SHT apparatus, i.e., that quenching at 300 °C/s is possible. The quenching specimens were then polished, and the diameter was measured in four directions, as shown in Figure 3a, to check the dissolution rate of the Al2O3 particles. However, using this method, only a small portion of the diameter of the Al2O3 particles may be exposed. For this reason, the largest diameter value found by repeating the polishing several times was set as the representative value of the Al2O3 particle dissolution rate, as shown in Figure 3b.
The quenching specimens were also analyzed using SEM to identify the reaction layer at the interface between the slag and Al2O3 particles. No compounds were observed at the interface in the specimens.

3. Results and Discussion

3.1. Dissolution Behavior of Al2O3 Particles according to Temperature and FeO Content in Slag

The SHT apparatus was used to assess the dissolution behavior of Al2O3 particles in slag with changing FeO content. The experiment temperatures were 1550, 1575, and 1600 °C, and the experiment times were 120, 240, and 360 s for each condition. Experiments were performed at least three times for each condition for reproducibility.
The variation in the diameter of Al2O3 particles in slags with different FeO contents at each temperature is shown in Figure 4. Slag0 and slag3 are based on previous papers [5,15]. In the case of slag3, it was not possible to conduct experiments at temperatures above 1550 °C due to the short circuit of B-type thermocouples at high temperatures. Also, in the case of slag0, the temperature deviation was increased by 50 °C, and hence there are no data at 1575 °C.
Figure 4a shows the dissolution behavior of Al2O3 particles at 1550 °C. The diameter of the Al2O3 particles decreased linearly with the dissolution time, and the dissolution rate increased as the FeO content in the slag increased. However, for slag3, the dissolution rate did not increase with the increasing FeO content. These experimental results were ascribed to the generation of the CA6 phase at the interface of Al2O3 particles and slag under the experimental conditions of slag3, which changed the dissolution process of the particles into an inter-compound chemical reaction [5]. Figure 4b shows the dissolution behavior of Al2O3 particles at 1575 °C. The results of the experiment at 1575 °C showed that the diameter of the Al2O3 particles decreased linearly with the dissolution time, and the dissolution rate increased with an increase in FeO content in the slag. Figure 4c shows the dissolution behavior of the Al2O3 particles at 1600 °C. The results of the experiment at 1600 °C showed that the diameter of the Al2O3 particles decreased linearly with the dissolution time, and the dissolution rate increased with an increase in FeO content in the slag. However, at 1600 °C in the slag2 experiment, the Al2O3 particles completely dissolved before 360 s. For this reason, additional experiments were conducted at 300 s for an accurate interpretation.
Table 3 lists the particle diameters of the Al2O3 particles for each condition.

3.2. Analysis of Slag/Al2O3 Particle Interface through SEM

As previously described, H. Um et al. [5] reported that for slag3, a CA6 phase was created at the interface of slag and Al2O3 particles, which changed the dissolution process of the Al2O3 particles. In addition, Park et al. [8] reported that a ring-shaped compound was formed along the particle/slag interface depending on the slag composition. For this reason, to better understand the dissolution behavior of Al2O3 particles in the experimental slag, the interface of slag and Al2O3 particles was analyzed using SEM and EDS. Figure 5 shows the results of SEM and EDS analysis for the interface of slag1, slag2, and Al2O3 particles under the condition of 1550 °C.
From Figure 5a,b, it can be seen that no compounds were formed at the interface of slag1, slag2, and Al2O3 particles. It was also confirmed that no compounds were formed at the interface at 1550, 1575, or 1600 °C. S. Yeo et al. [15] and Taira et al. [24] reported that the dissolution rate of Al2O3 in CaO-SiO2-Al2O3 slag was controlled by diffusion in the boundary layer. Also, H. Um et al. [5] reported that the dissolution of Al2O3 particles in CaO-SiO2-Al2O3-FexO slag was controlled by diffusion in the boundary layer of slag and Al2O3 particles if no compounds were generated at the interface of slag and Al2O3 particles. Furthermore, in the present study, it can be seen from Figure 5c,d that the concentration of Al decreased linearly along the boundary layer from Al2O3 particles to slag. On the other hand, the concentrations of Ca, Si, and Fe were found to have increased linearly. For this reason, it was determined that Al2O3 in the boundary layer exhibited diffusion behavior.
Therefore, no compounds were formed at the interface of Al2O3 particles and slag under the conditions of this study, and based on the EDS results of the interface, it is believed that for the dissolution mechanism of Al2O3 particles in slag1 and slag2 it was only necessary to consider the behavior by diffusion in the boundary layer.

3.3. Dissolution Mechanism of Al2O3

Solid Al2O3 particles can be dissolved by liquid slag through the following process [5].
Al 2 O 3 s =   Al 2 O 3 l Chemical   kinetics   of   the   reaction   at   the   interface
Al 2 O 3 l = Al 2 O 3    Liquid-phase   mass   transfer
In other words, the dissolution of solid Al2O3 particles in liquid slag can be controlled by a chemical reaction or liquid-phase mass transfer [8,24,25,26]. It is understood that the dissolution of Al2O3 particles in liquid slag is controlled by liquid-phase mass transfer in the boundary layer unless a compound is generated at the interface of Al2O3 particles and slag [9,19,25,27,28,29,30]. As noted above, no compounds were generated at the interface of Al2O3 particles and slag under the present experimental conditions. Therefore, the rate-controlling step of Al2O3 particle dissolution in this experiment can be interpreted as liquid-phase mass transfer at the boundary layer.
If the rate-controlling step in the dissolution of Al2O3 in slag is liquid-phase mass transfer in the boundary layer, then the relationship between the mass transfer flux and mass transfer coefficient of Al2O3 particle dissolution can be expressed by the mass transfer equation as follows [31]:
J = k C i C b
where J is the mass transfer flux; k is the mass transfer coefficient in the slag; Ci and Cb are the Al2O3 content at the interface and in bulk slag; and ( C i C b ) is the driving force of the dissolution.
If the Al2O3 particles being dissolved are spherical, then Equation (3) can be transformed into the following dissolution rate equation [31]:
d r d t = k C i C b M / ρ
where r is the radius of the Al2O3 particles; dr/dt is the dissolution rate; M is the molecular weight of Al2O3; and ρ is the slag density.
The dissolution rate can be calculated from the experimental data, as shown in Figure 6, and the slag density and driving force of the dissolution can be obtained using FactSage7.3TM; hence, the mass transfer coefficient according to temperature and FeO content can be derived using Equation (4). The physical properties and mass transfer coefficients for each temperature and FeO content are summarized in Table 4.
The mass transfer coefficient increases with the FeO content in the slag and with increasing melting temperature.

3.4. Activation Energy

The activation energy of mass transfer—the dissolution mechanism of Al2O3—can be quantitatively measured. In this study, the dissolution rate of Al2O3 particles was measured with temperature as a variable, and the mass transfer coefficient was calculated accordingly. By graphing the reciprocal of the mass transfer coefficient and temperature using the Arrhenius equation, which is expressed as follows, the activation energy for mass transfer can be derived [32]:
k = k 0 exp E k R T
where k is the mass transfer coefficient; k0 is the pre-exponential constant; R is the universal gas constant; T is the absolute temperature; and Ek is the activation energy of mass transfer.
To explain how Ek is derived, substituting logarithms into the above expression, we can express it as a function of ln(k) and temperature, as follows:
l n k = E k R 1 T + l n   k 0
This expression tells us that the slope of the ln(k)-1/T graph is E k R . Therefore, multiplying this value by −R gives the activation energy for mass transfer. Cho et al. calculated the mass transfer coefficient by measuring the dissolution behavior of Al2O3 in CaO-SiO2-Al2O3 slag with temperature as a variable, and derived the Ek of Al2O3 dissolution using the above method [31]. In this study, Ek was also calculated through the above process and is shown in Figure 7.
The Ek of slag1 was calculated to be 159 kJ/mole and the Ek of slag2 was calculated to be 182 kJ/mole. In addition, the Ek of slag with FeO calculated in this study and the Ek values of slag without FeO calculated in other studies are summarized in Table 5.
From Table 5, the Ek of the experimental slag with FeO is lower than that of the slag without FeO, indicating that the Al2O3 dissolution in the slag with FeO is faster compared to that in the slag without FeO. In other words, when FeO is included in the slag, the mass transfer of Al2O3 particles is relatively easier and the dissolution of Al2O3 particles can be faster. This means that, as mentioned at the introduction, when slag is used to remove Al2O3 inclusions, electric furnace slag containing a large amount of FeO will be better at removing Al2O3 inclusions than blast furnace slag containing a trace amount of FeO.

3.5. Increased Dissolution Rate of Al2O3 by Increasing FeO Content

We can interpret the increase in the dissolution rate of Al2O3 particles with increasing FeO content in the slag in terms of viscosity and the driving force of dissolution. The viscosity of slag0, 1, and 2 at the experimental temperature was obtained using FactSage7.3TM. Additionally, a phase diagram was drawn using FactSage7.3TM and is shown in Figure 8, and the driving force for dissolution of Al2O3 particles was obtained using this. We used FactPS, FT oxid, and FS stel databases of FactSage7.3TM to draw the state diagram. The results are summarized in Table 6.
From Figure 8, it can be seen that the liquid-phase area increased with increasing FeO content and temperature, which also increased the driving force of dissolution.
From Table 6, it can be seen that the viscosity of the slag decreased and Al2O3 driving force of dissolution increased as the FeO content increased.
In summary, as the FeO content in the slag increased, the slag viscosity decreased and the driving force of Al2O3 dissolution increased. In addition, in this study, the mechanism of Al2O3 dissolution was mass transfer. Therefore, the slag viscosity, the driving force of dissolution of Al2O3 particle, and the dissolution rate of Al2O3 particle have the following relationship [34,35]:
log r   log ( Δ C η )
where r is the dissolution rate of Al2O3; ΔC is the driving force for the dissolution of Al2O3; and η is the slag viscosity.
To summarize this equation and the above relationship, as the FeO content increased, the viscosity decreased, and the driving force of Al2O3 particle dissolution increased, increasing the dissolution rate of Al2O3 dissolution. In fact, the result of this experiment shows that the dissolution rate increased as the FeO content increased. In other words, it can be seen that the results of the experiment fit well with the relationship of Equation (7). In addition, FeO was added and the viscosity decreased, so diffusion occurred more easily. As a result, the liquid-phase mass transfer was accelerated, and the mass transfer coefficient increased.

4. Conclusions

In this study, the dissolution behavior of Al2O3 was studied by varying the FeO content in the slag from 10 to 20 wt%, and temperatures of 1550 °C, 1575 °C, and 1600 °C. The results were as follows:
(1)
The dissolution rate increased linearly as the FeO content of the slag increased from 0 to 20 wt% and the dissolution temperature increased from 1550 to 1600 °C.
(2)
Through an SEM and EDS analysis, it was observed that no compound was formed at the interface of Al2O3 particles and slag. In addition, it was observed that the concentration of Al in the boundary layer decreased linearly as it moved from Al2O3 particles to slag. Therefore, the rate step of Al2O3 particle dissolution is interpreted as liquid-phase mass transfer.
(3)
The mass transfer coefficient was obtained using the dissolution rate equation. The mass transfer coefficient increased with increasing FeO content in the slag and increasing dissolution temperature.
(4)
The mass transfer coefficient was plotted in a graph as a function of temperature, and the Ek values of slag1 and slag2 (159 and 182 kJ/mole, respectively) were found using the Arrhenius equation.
(5)
The Ek of Al2O3 mass transfer in slag containing FeO in this study was lower than the Ek of slag without FeO.
(6)
As the FeO content in the slag increased, the viscosity decreased and Al2O3 dissolution driving force increased, resulting in an increase in the dissolution rate. Additionally, as viscosity decreased, liquid-mass transfer occurred more easily and the mass transfer coefficient increased.

Author Contributions

Conceptualization, T.K., H.U. and Y.C.; Methodology, T.K., H.U.; Validation, T.K.; Investigation, T.K.; Writing – original draft, T.K.; Writing – review & editing, H.U. and Y.C.; Supervision, H.U. and Y.C.; Project administration, Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Korea Evaluation Institute of Industrial Technology (KEIT) grant (Grant number 00262191, 00262711), and by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant (Grant number 20212010100060).

Data Availability Statement

No data available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fan, Z.; Friedmann, S.J. Low-carbon production of iron and steel: Technology options, economic assessment, and policy. Joule 2021, 5, 829–862. [Google Scholar] [CrossRef]
  2. Lee, B.; Sohn, I. Review of Innovative Energy Savings Technology for the Electric Arc Furnace. JOM 2014, 66, 1581–1594. [Google Scholar] [CrossRef]
  3. Park, J.H.; Todoroki, H. Control of MgO·Al2O3 spinel inclusions in stainless steels. ISIJ Int. 2010, 50, 1333–1346. [Google Scholar] [CrossRef]
  4. Dimitrov, S.; Weyl, A.; Janke, D. Control of the aluminium-oxygen reaction in pure iron melts. Steel Res. 1995, 66, 3–7. [Google Scholar] [CrossRef]
  5. Um, H.; Yeo, S.; Kang, Y.-B.; Chung, Y. The effect of FexO content on dissolution behavior of an alumina inclusion in CaO–Al2O3–SiO2–FexO slag by a single hot thermocouple technique. Ceram. Int. 2022, 48, 35301–35309. [Google Scholar] [CrossRef]
  6. Jung, I.-H.; Decterov, S.A.; Pelton, A.D. Computer applications of thermodynamic databases to inclusion engineering. ISIJ Int. 2004, 44, 527–536. [Google Scholar] [CrossRef]
  7. Holappa, L.; Hämäläinen, M.; Liukkonen, M.; Lind, M. Thermodynamic examination of inclusion modification and precipitation from calcium treatment to solidified steel. Ironmak. Steelmak. 2003, 30, 111–115. [Google Scholar] [CrossRef]
  8. Park, J.-H.; Jung, I.-H.; Hae-Geon, L.E.E. Dissolution behavior of Al2O3 and MgO inclusions in the CaO-Al2O3-SiO2 slags: Formation of ring-like structure of MgAl2O4 and Ca2SiO4 around MgO inclusions. ISIJ Int. 2006, 46, 1626–1634. [Google Scholar] [CrossRef]
  9. Sridhar, S.; Cramb, A.W. Kinetics of Al2O3 dissolution in CaO-MgO-SiO2-Al2O3 slags: In situ observations and analysis. Metall. Mater. Trans. B Process Metall. Mater. Process. Sci. 2000, 31, 406–410. [Google Scholar] [CrossRef]
  10. Shu, Q.; Zhang, X.; Wang, Y.; Li, J.; Chou, K. Effect of Na2O on dissolution rate of alumina in CaO-Al2O3-MgO-SiO2 slag. In Proceedings of the 6th International Congress on the Science and Technology of Steelmaking, ICS, Beijing, China, 12–14 May 2015; pp. 606–609. [Google Scholar]
  11. Yi, K.W.; Tse, C.; Park, J.-H.; Valdez, M.; Cramb, A.W.; Sridhar, S. Determination of dissolution time of Al2O3 and MgO inclusions in synthetic Al2O3-CaO-MgO slags. Scand. J. Metall. 2003, 32, 177–184. [Google Scholar] [CrossRef]
  12. Ren, C.; Zhang, L.; Zhang, J.; Wu, S.; Zhu, P.; Ren, Y. In Situ Observation of the Dissolution of Al2O3 Particles in CaO-Al2O3-SiO2 Slags. Metall. Mater. Trans. B Process Metall. Mater. Process. Sci. 2021, 52, 3288–3301. [Google Scholar] [CrossRef]
  13. Holappa, L.; Kekkonen, M.; Louhenkilpi, S.; Hagemann, R.; Schröder, C.; Scheller, P. Active tundish slag. Steel Res. Int. 2013, 84, 638–648. [Google Scholar] [CrossRef]
  14. Park, Y.-J.; Cho, Y.-M.; Cha, W.-Y.; Kang, Y.-B. Dissolution kinetics of alumina in molten CaO–Al2O3–FetO–MgO–SiO2 oxide representing the RH slag in steelmaking process. J. Am. Ceram. Soc. 2020, 103, 2210–2224. [Google Scholar] [CrossRef]
  15. Yeo, S.; Um, H.; Chung, Y. The Effect of Alumina Activity on Dissolution Behavior of Alumina Particles in CaO–Al2O3–SiO2 Slags. Metall. Mater. Trans. B Process Metall. Mater. Process. Sci. 2021, 52, 3938–3945. [Google Scholar] [CrossRef]
  16. Zhang, S.; Rezaie, H.R.; Sarpoolaky, H.; Lee, W.E. Alumina dissolution into silicate slag. J. Am. Ceram. Soc. 2000, 83, 897–903. [Google Scholar] [CrossRef]
  17. Valdez, M.; Prapakorn, K.; Cramb, A.W.; Seetharaman, S. A study of the dissolution of Al2O3, MgO and MgAl2O4 particles in a CaO-Al2O3-SiO2 slag. Steel Res. 2001, 72, 291–297. [Google Scholar] [CrossRef]
  18. Monaghan, B.J.; Chen, L.; Sorbe, J. Comparative study of oxide inclusion dissolution in CaO-SiO2-Al2O3 slag. Ironmak. Steelmak. 2005, 32, 258–264. [Google Scholar] [CrossRef]
  19. Chen, G.; He, S.; Wang, Q. Dissolution behavior of Al2O3 into tundish slag for high-al steel. J. Mater. Res. Technol. 2020, 9, 11311–11318. [Google Scholar] [CrossRef]
  20. Shi, G.-Y.; Zhang, T.-A.; Dou, Z.-H.; Niu, L.-P. Dissolution behavior of Al2O3 inclusions in CaO-Al2O3 based slag representing aluminothermic reduction slag. Crystals 2020, 10, 1061. [Google Scholar] [CrossRef]
  21. Odenthal, H.-J.; Kemminger, A.; Krause, F.; Sankowski, L.; Uebber, N.; Vogl, N. Review on Modeling and Simulation of the Electric Arc Furnace (EAF). Steel Res. Int. 2018, 89, 1700098. [Google Scholar] [CrossRef]
  22. Lee, S.; Chung, Y. The effect of C content in MgO–C on dissolution behavior in CaO–SiO2–Al2O3 slag. Ceram. Int. 2022, 48, 26984–26991. [Google Scholar] [CrossRef]
  23. Kim, Y.; Kashiwaya, Y.; Chung, Y. Effect of varying Al2O3 contents of CaO–Al2O3–SiO2 slags on lumped MgO dissolution. Ceram. Int. 2020, 46, 6205–6211. [Google Scholar] [CrossRef]
  24. Taira, S.; Nakashima, K.; Mori, K. Kinetic Behavior of Dissolution of Sintered Alumina Into CaO-SiO2&Al2O3 Slags. ISIJ Int. 1993, 33, 116–123. [Google Scholar] [CrossRef]
  25. Choi, J.-Y.; Lee, H.-G.; Kim, J.-S. Dissolution rate of Al2O3 into molten CaO-SiO2-Al2O3 slags. ISIJ Int. 2002, 42, 852–860. [Google Scholar] [CrossRef]
  26. Samaddar, B.N.; Kingery, W.D.; Cooper, A.R. Dissolution in Ceramic Systems: 11, Dissolution of Aluminu, Mullite, Anorthite, and Silica in a Calcium-Aluminum-Silicate Slag. J. Am. Ceram. Soc. 1964, 47, 249–254. [Google Scholar] [CrossRef]
  27. Oishi, Y.; Cooper, A.R.; Kingery, W.D. Dissolution in Ceramic Systems: III, Boundary Layer Concentration Gradients. J. Am. Ceram. Soc. 1965, 48, 88–95. [Google Scholar] [CrossRef]
  28. Bui, A.-H.; Ha, H.-M.; Kang, Y.-B.; Chung, I.-S.; Lee, H.-G. Dissolution behavior of alumina in mold fluxes for steel continuous casting. Met. Mater. Int. 2005, 11, 183–190. [Google Scholar] [CrossRef]
  29. Yan, P.; Webler, B.A.; Pistorius, P.C.; Fruehan, R.J. Nature of MgO and Al2O3 Dissolution in Metallurgical Slags. Metall. Mater. Trans. B Process Metall. Mater. Process. Sci. 2015, 46, 2414–2418. [Google Scholar] [CrossRef]
  30. Sharma, M.; Mu, W.; Dogan, N. In Situ Observation of Dissolution of Oxide Inclusions in Steelmaking Slags. JOM 2018, 70, 1220–1224. [Google Scholar] [CrossRef]
  31. Cho, W.D.; Fan, P. Diffusional Dissolution of Alumina in Various Steelmaking Slags. ISIJ Int. 2004, 44, 229–234. [Google Scholar] [CrossRef]
  32. Petrucci, R.H.; Harwood, W.S. General Chemistry, 6th ed.; Macmillan: New York, NY, USA, 1993; p. 535. [Google Scholar]
  33. Liu, Y.Q.; Wang, L.J.; Chou, K.C. Dissolution behavior of Al2O3 in refining slags containing Ce2O3. ISIJ Int. 2014, 54, 728–733. [Google Scholar] [CrossRef]
  34. Park, J.S.; Park, J.H. Effect of Physicochemical Properties of Slag and Flux on the Removal Rate of Oxide Inclusion from Molten Steel. Metall. Mater. Trans. B Process Metall. Mater. Process. Sci. 2016, 47, 3225–3230. [Google Scholar] [CrossRef]
  35. Valdez, M.; Shannon, G.S.; Sridhar, S. The ability of slags to absorb solid oxide inclusions. ISIJ Int. 2006, 46, 450–457. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of single hot thermocouple apparatus.
Figure 1. Schematic diagram of single hot thermocouple apparatus.
Metals 13 01702 g001
Figure 2. Experimental procedure of Al2O3 particle dissolution.
Figure 2. Experimental procedure of Al2O3 particle dissolution.
Metals 13 01702 g002
Figure 3. Measurement of alumina diameter in opaque slag: (a) measurement of Al2O3 particle in polished quenching specimens; (b) the largest diameter of Al2O3 particles found through multiple polishings.
Figure 3. Measurement of alumina diameter in opaque slag: (a) measurement of Al2O3 particle in polished quenching specimens; (b) the largest diameter of Al2O3 particles found through multiple polishings.
Metals 13 01702 g003
Figure 4. Al2O3 particle dissolution behavior according to FeO content and temperature: (a) 1550 °C; (b) 1575 °C; (c) 1600 °C. Data by Yeo et al. refer to Ref. [15]; Data by Um et al. refer to Ref. [5].
Figure 4. Al2O3 particle dissolution behavior according to FeO content and temperature: (a) 1550 °C; (b) 1575 °C; (c) 1600 °C. Data by Yeo et al. refer to Ref. [15]; Data by Um et al. refer to Ref. [5].
Metals 13 01702 g004
Figure 5. Cross-sectional image of Al2O3 particles from SEM: (a) Al2O3 particle in slag1; (b) Al2O3 particle in slag2; and spectrum of EDS line scanning of Al, Si, Ca, and Fe at the Al2O3 particles/slag interface; (c) Al2O3 particle/slag1; (d) Al2O3 particle/slag2.
Figure 5. Cross-sectional image of Al2O3 particles from SEM: (a) Al2O3 particle in slag1; (b) Al2O3 particle in slag2; and spectrum of EDS line scanning of Al, Si, Ca, and Fe at the Al2O3 particles/slag interface; (c) Al2O3 particle/slag1; (d) Al2O3 particle/slag2.
Metals 13 01702 g005aMetals 13 01702 g005b
Figure 6. Dissolution rate by temperature and FeO content.
Figure 6. Dissolution rate by temperature and FeO content.
Metals 13 01702 g006
Figure 7. Arrhenius plot of the mass transfer coefficient.
Figure 7. Arrhenius plot of the mass transfer coefficient.
Metals 13 01702 g007
Figure 8. CaO-Al2O3-SiO2-FeO phase diagram and composition of each slag at the experiment temperature: (a) slag0; (b) slag1; and (c) slag2.
Figure 8. CaO-Al2O3-SiO2-FeO phase diagram and composition of each slag at the experiment temperature: (a) slag0; (b) slag1; and (c) slag2.
Metals 13 01702 g008aMetals 13 01702 g008b
Table 1. Chemical compositions and diameter of Al2O3 particles.
Table 1. Chemical compositions and diameter of Al2O3 particles.
TypeSourceDiameterWeightConcentration
Al2O3 (%)Other (%)
Alumina sphereGoodFellow500 ± 2.5 μm0.25 ± 0.05 mg99.90.1
Table 2. Chemical compositions of slag (wt%).
Table 2. Chemical compositions of slag (wt%).
HeadingsCaOSiO2Al2O3FeOBasicityReferences
Slag047.547.5501[15]
Slag143.742.64.69.11
Slag237.738.44.419.51
Slag332.532.55301[5]
Table 3. Diameter of Al2O3 particles according to FeO content and temperature (μm).
Table 3. Diameter of Al2O3 particles according to FeO content and temperature (μm).
Temperature (°C)Slag120 s240 s360 sReferences
15500465420393[15]
1408360269
2353251122
3377283122[5]
15751407321248
232419485
16000447385317[15]
1374240187
231910555 (300 s)
Table 4. Physical properties and mass transfer coefficient according to temperature and FeO content.
Table 4. Physical properties and mass transfer coefficient according to temperature and FeO content.
Driving Force of Dissolution Δ C (mole/m3) *FactSage7.3TMSlag Density ρ
(kg/m3)
Dissolution Rate d r d t (cm/s)Mass Transfer Coefficient k
(cm/s)
References
Slag0 1550 °C10,74526602.96 × 10−57.19 × 10−8[15]
Slag0 1600 °C11,39926475.10 × 10−51.16 × 10−7
Slag1 1550 °C10,88527966.18 × 10−51.56 × 10−7Present study
Slag1 1575 °C11,58127897.01 × 10−51.66 × 10−7
Slag1 1600 °C11,84627828.95 × 10−52.06 × 10−7
Slag2 1550 °C11,05029431.03 × 10−42.69 × 10−7Present study
Slag2 1575 °C11,60029351.15 × 10−42.84 × 10−7
Slag2 1600 °C11,92529281.54 × 10−43.71 × 10−7
Table 5. Comparison of Al2O3 dissolution Ek according to slag composition.
Table 5. Comparison of Al2O3 dissolution Ek according to slag composition.
SlagChemical Composition (wt%)Ek (kJ/mole)References
CaOSiO2Al2O3MgOCe2O3FeO
047.547.55.0000304[15]
142.542.55.00010159Present study
237.537.55.00020182
445.010.045.0000445[31]
535.030.035.0000334
645.04.537.510.030292[33] 1
745.04.535.510.050347
845.04.532.510.080249
1 Cylindrical Al2O3 rotated at 200 rpm.
Table 6. Slag viscosity and Al2O3 particle driving force of dissolution by temperature and FeO content.
Table 6. Slag viscosity and Al2O3 particle driving force of dissolution by temperature and FeO content.
HeadingsSlag Viscosity
(Pa·s)
*FactSage7.3TM
Driving Force of Dissolution Δ C (mole/m3) *FactSage7.3TMReferences
Slag0 1550 °C3.23510,745[15]
Slag0 1600 °C2.20211,399
Slag1 1550 °C0.15910,885
Slag1 1575 °C0.14011,581Present study
Slag1 1600 °C0.12411,846
Slag2 1550 °C0.09811,050
Slag2 1575 °C0.08711,600Present study
Slag2 1600 °C0.07811,925
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

Kwack, T.; Um, H.; Chung, Y. Activation Energy of Alumina Dissolution in FeO-Bearing Slags. Metals 2023, 13, 1702. https://doi.org/10.3390/met13101702

AMA Style

Kwack T, Um H, Chung Y. Activation Energy of Alumina Dissolution in FeO-Bearing Slags. Metals. 2023; 13(10):1702. https://doi.org/10.3390/met13101702

Chicago/Turabian Style

Kwack, Taejun, Hyungsic Um, and Yongsug Chung. 2023. "Activation Energy of Alumina Dissolution in FeO-Bearing Slags" Metals 13, no. 10: 1702. https://doi.org/10.3390/met13101702

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