Next Article in Journal
Aluminum Alloys and Aluminum-Based Matrix Composites
Next Article in Special Issue
Electrochemical Behavior of Niobium Oxide and Titanium Oxide in NaF–Na3AlF6 Molten Salt
Previous Article in Journal
Hydrogen Production Properties of Aluminum–Magnesium Alloy Presenting β-Phase Al3Mg2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Substituting FeO with CaO on the Rheological and Surface Properties of Silicate Melts

1
Key Laboratory for Ecological Metallurgy of Multimetallic Ores (Ministry of Education), Shenyang 110819, China
2
School of Metallurgy, Northeastern University, Shenyang 110819, China
*
Author to whom correspondence should be addressed.
Metals 2023, 13(11), 1869; https://doi.org/10.3390/met13111869
Submission received: 18 October 2023 / Revised: 7 November 2023 / Accepted: 8 November 2023 / Published: 9 November 2023
(This article belongs to the Special Issue Advances in Slag Metallurgy—Second Edition)

Abstract

:
A comprehensive understanding of the structural impact of composition is crucial in designing converter slag to optimize its rheological and surface properties during the smelting process. In this study, glassy CaO-SiO2-FexO samples with varying CaO/FexO ratios were prepared to simulate the slag in the initial stage of converter melting. The viscosity and surface tension of the slag at 1300–1600 °C were measured, and the microscopic essence of physical properties was further analyzed using Raman spectroscopy technology. The findings reveal that as CaO replaces FeO, [SiO4]-tetrahedra gradually depolymerize from Q4(Si) to Q0(Si), while [FeO6]-octahedra gradually transform into [FeO4]-tetrahedra, resulting in a decrease in the degree of polymerization of the slag. The slag with a lower degree of polymerization exhibits reduced activation energy of viscous flow and increased surface tension. Therefore, it is of great significance to appropriately control the CaO/FexO ratio in the early stage of smelting to improve the rheological and surface properties of the slag.

1. Introduction

Converter steelmaking is a pivotal step in the contemporary steel material preparation process, playing a crucial role in subsequent refining, solidification, and final quality control of liquid steel. Steelmaking is essentially a slagging process wherein the slag assumes an important metallurgical function by enhancing the steel quality and smelting efficiency [1,2]. As converter steelmaking progresses towards high-quality, high-efficiency, environmentally friendly, and safe practices, there is an increasing demand for the metallurgical function of slag that closely correlates with its physical properties. Therefore, it becomes imperative to establish a comprehensive understanding of the relationship between its physical properties, metallurgical function, and chemical composition to effectively enhance its metallurgical performance.
With the advancement of research, there has been a growing recognition of the essence of the slag microstructure, leading to extensive exploration of its correlation with macroscopic physical properties and metallurgical functionalities. In investigations on slag viscosity, Gi et al. [3] observed that the substitution with larger alkali metal cations, e.g., Li+, Na+, and K+, increased the viscosity of the melts and enhanced the degree of polymerization of the Q3(Si) structural units. Wen et al. [4] found that the addition of SiO2 leads to the increase of the slag viscosity and the activation energy increases. According to deconvolution results of XPS, as SiO2 content in glassy slag increases, the number of bridging oxygens increases, indicating a more polymerized structure and a larger viscosity.
In the investigation of surface tension, Gao et al. [5] discovered a significant increase in the surface tension of CaO-SiO2-Na2O-CaF2 slag with an increasing CaO/SiO2 ratio, which could be attributed to the depolymerization of the slag. Sukenaga et al. [6] investigated the effect of CaO/SiO2 on the surface tension of CaO-SiO2-Al2O3-MgO slag within a temperature ranging from 1723 to 1823 K and found that with the increase in the CaO/SiO2 ratio from 1.1 to 1.7, a large amount of unsaturated non-bridging oxygen was formed on the melt surface, resulting in an increase of surface tension. The surface tension is closely related to its thermodynamic properties. Soledade et al. [7] described about the significance of the Butler equation for the examination of the surface tension of the liquid mixture. At present, surface tension models are mostly established through this model.
Current research mainly focuses on investigating the impact of changes in the CaO/SiO2 ratio on the physical properties of slag. However, in the converter smelting process, continuous dissolution of limestone leads to ongoing substitution of FeO by CaO in the slag [8]. CaO/FexO is also an important factor affecting the physical properties. Therefore, it is essential to systematically analyze the structural behavior of ions to explore the transformation of physical properties. The present work investigated the viscosity, surface tension, and structure of a simplified CaO-SiO2-FexO slag to reveal the transformation of physical properties from a microscopic perspective. The research findings contribute to a deeper comprehension of the micro-level transformation mechanism governing macroscopic physical properties and metallurgical behavior of converter slag during the smelting process, thereby providing guidance for the refined design of slag composition and precise regulation of the smelting process.

2. Materials and Methods

2.1. Preparation of Slag Samples

Reagent-grade CaO, SiO2, and FeC2O4∙2H2O powders were utilized to synthesize the samples. When the temperature exceeds 850 °C, FeC2O4·2H2O can decompose into FeO. To investigate the effects of replacing FeO with CaO on slag properties and microstructure, six experimental slags were designed with varying CaO content (5%, 15%, 25%, 35%, 40%, and 50%). The preparation process of quenched slag has been described in our previous research [9]. The quenched slag was subjected to X-ray fluorescence spectroscopy (XRF) and X-ray diffraction (XRD) analysis to determine whether a glassy sample with satisfactory chemical composition was prepared. Table 1 presents the chemical composition of the slag, wherein CaO and SiO2 contents are consistent with the design composition. Some FeO is oxidized to Fe2O3, and the Fe2+/TFe ratio is maintained at around 0.7. Figure 1 shows a typical XRD pattern of the slag. All the XRD profiles only showed a broad peak around the diffraction angle 2θ of 30°, which belongs to the typical glass phase characteristics.

2.2. Performance Testing

The viscosity and surface tension of CaO-SiO2-FexO slag were measured using the rotating cylinder method and the pulling tube method, respectively. Figure 2 shows the experimental equipment, wherein temperature control is achieved through a microcomputer program with a precision of ±0.5 °C. Surface tension tests were conducted to obtain the average value from six measurements of the surface tension at that temperature. Meanwhile, three viscosity tests were performed, and their average value represented slag viscosity.

2.3. Microstructure Detection

The microstructure of quenched slag is analyzed using a LabRAM HR800 Raman spectrometer equipped with a CCD detector. Raman spectra were recorded using an Ar+ laser with an excitation wavelength of 488 nm as the light source. The scanning range of the spectrum is 200–2000 cm−1, and the spectral resolution is 0.65 cm−1. The testing curve was imported into Origin 8.5 software, and the smoothing and baseline correction were performed to remove noise and fluorescence effects, respectively. Referring to the relevant literature, an appropriate frequency range was selected to deconvolute by Gaussian function.

3. Results

3.1. Viscosity

The viscosity curves of CaO-SiO2-FexO slag with different temperatures and CaO/FexO ratios are presented in Figure 3. In Figure 3a, the slag viscosity generally decreases with increasing temperature and substitution of CaO for FeO. Notably, the addition of CaO causes a significant decrease in the viscosities of CSF1-3 slags, indicating that CaO exerts a stronger influence on slag viscosity compared to FeO.
The relationship between viscosity and temperature is commonly described using the Arrhenius formula, where lnη and T−1 are linearly correlated. The linear regression analysis is shown in Figure 3b. The fitting coefficients of each slag are greater than 0.95, demonstrating a strong linear association between lnη and T−1. From this, the activation energies of the viscous flow of CSF1-6 slags can be determined as 160, 139, 133, 115, 110, and 109 kJ·mol−1, respectively.
The activation energy of viscous flow in metallurgical melts can be defined as the total energy required for particles to form holes and migrate between holes. As CaO replaces FeO, the decreasing activation energy at 1300–1600 °C indicates an increase in the number of holes in the melt and an enhancement of migration ability, thereby causing a decrease in slag viscosity. However, when the addition of CaO exceeds 35%, there is only a relatively small reduction in the activation energy observed, suggesting that it reaches a lower state.

3.2. Surface Tension

Figure 4 shows the surface tension curves of the slags under different CaO/FexO conditions as a function of temperature. It can be seen that the surface tension of the slag gradually increases with the substitution of CaO for FeO at the same temperature; as the temperature increases, the surface tension of each slag shows a linear decrease, the slope of which represents the temperature coefficient of surface tension. The temperature coefficients are sequentially obtained as −3.16, −3.14, −2.9, −2.44, and −1.8, respectively. As CaO content increases, the absolute value of the temperature coefficient shows a decreasing trend, and the decrease amplitude gradually increases. This indicates that CaO weakens the dependence of surface tension on temperature, and the higher content of CaO results in even weaker reliance on changes in temperature.

3.3. Melt Structure

Figure 5 shows the Raman spectra of CaO-SiO2-FexO slag in the frequency range of 400–1200 cm−1. The vibration peaks in the low-frequency region below 500 cm−1 are believed to be associated with the deformation and vibration of tetrahedron [10,11], and no further analysis will be conducted. The intermediate frequency range consists of four vibration peaks near 535, 580, 670, and 710 cm−1. The vibration peaks near 535 and 710 cm−1 correspond to the bending vibrations of Si-O and Si-O-Si bonds [12,13], respectively, while the vibration peaks near 580 and 670 cm−1 represent [FeO6]-octahedra and [FeO4]-tetrahedra [14,15], respectively. As CaO replaces FeO, there is a gradual decrease in relative intensity observed for these vibration peaks at around 580 and 670 cm−1, which is consistent with the decreasing trend of FeO content from 55 wt.% to 10 wt.%.
The high-frequency region consists of six characteristic peaks, of which the peaks near 855, 927, 960, 1060, and 1130 cm−1 represent the [SiO4]-tetrahedra with bridging oxygen numbers of 0, 1, 2, 3, and 4 [11,16,17], respectively, and the peak near 1000 cm−1 is related to the vibration of Si-O-Fe bonds [18]. The Raman spectra are deconvoluted according to the above assignments, and the results are shown in Figure 6.
The relative area fraction of structural units is calculated to clearly analyze the changes in slag structure after CaO replacing FeO, as illustrated in Figure 7. With the increase of CaO, Q0(Si) content increases while Q3(Si) and Q4(Si) content decreases. Meanwhile, Q1(Si) and Q2(Si) exhibit fluctuations but show an overall decreasing trend. This can be attributed to the dissociation of O2− from both CaO and FeO, resulting in the gradual depolymerization of [SiO4]-tetrahedra by cutting off Si-O-Si bonds. The depolymerization process can be represented by Equation (1), which is consistent with the depolymerization mechanism proposed by Mysen et al. [19].
Metals 13 01869 i001
In addition, it was noted that [FeO4]-tetrahedra gradually increased while [FeO6]-octahedra decreased. This phenomenon is particularly pronounced at higher levels of CaO content. The dissociated O2− from CaO preferentially combines with Si4+ to form the Si-O bonds due to the stronger energy compared to Fe-O bonds, resulting in the depolymerization of the aforementioned [SiO4]-tetrahedra. As O2− further increases, excessive O2− combines with Fe3+ to form [FeO4]-tetrahedra, causing the increase of [FeO4]/[FeO6]. Similar results were reported by Rüssel and Wiedenroth [14], where an increase in basicity is beneficial for the stability of [FeO4]-tetrahedra. Despite enhancing the networking role of Fe3+, the slag still exhibits the depolymerization of [SiO4]-tetrahedra and the degree of polymerization of the slag is reduced.

4. Discussion

4.1. Effect of CaO-FeO Substitution on Rheological Property

To elucidate the relationship between slag viscosity and its structure, a linear fit between the activation energy of viscous flow and the structural parameters is plotted in Figure 8. The activation energy is inversely proportional to the molar fractions of simple [SiO4]-tetrahedra, while it is directly proportional to the molar fractions of complex [SiO4]-tetrahedra. This indicates that the activation energy of viscous flow is smaller in the melts with lower polymerization degrees, which is consistent with the study by Yang et al. [20].
The viscosity variation of CaO-SiO2-FexO slag can be microscopically analyzed, as shown in Figure 9. When the CaO addition is less than 35%, the substitution of FeO by CaO leads to preferential destruction of Si-O-Si bonds by dissociated O2− due to its strong affinity with Si4+ [21]. Consequently, significant depolymerization is observed in highly polymerized Q4(Si) and Q3(Si); complete depolymerization of Q4(Si) occurs when CaO content reaches 25%. Since Fe3+ has limited O2− around it, some Fe3+ ions act as network modifiers. Therefore, the significant decrease in viscosity with increasing CaO at this stage can be attributed to the depolymerization of highly polymerized [SiO4]-tetrahedra in the slag.
When CaO content exceeds 35%, O2− destroys the Si-O-Si bonds in Q3(Si) and Q2(Si), depolymerizing into simpler Q1(Si) and Q0(Si). Furthermore, an increase in O2− concentration promotes the formation of [FeO4]-tetrahedra by Fe3+, resulting in a reduction in free Fe3+ content [14]. However, the slag remains in a state of dissociation overall. Therefore, the slag viscosity gradually decreases, but the reduction amplitude of viscosity decreases due to the increase of [FeO4]/[FeO6].

4.2. Effect of CaO-FeO Substitution on Surface Property

For oxide melts, ions are primarily bound together by electrostatic force, and strong electrostatic forces can lead to significant surface tension. Generally, the electrostatic force is inversely proportional to the distance between ions [22]. Based on this relationship, the reason for the decrease in surface tension of slag with increasing temperature can be explained.
On the one hand, the irregular thermal motion of ions increases with temperature, resulting in a larger distance between ions, which weakens the interaction force between ions. On the other hand, as the temperature increases, simple [SiO4]-tetrahedra increases while complex [SiO4]-tetrahedra decreases, resulting in a reduction in the average ion radius of the slag. According to Einstein’s equation, both elevated temperature and decreased ion radius can increase the average displacement. This implies a decrease in activation energy required for ion motion and intensifies the irregular thermal motion of ions, causing larger interionic distances and reduced slag surface tension.
The Butler equation describes the relationship between surface tension, temperature, and activity, as follows in Equation (2) [23].
σ = σ i + R T A i ln a i S a i B
The surface tension of slag is influenced by its microstructure, so it is assumed that the Butler equation is a function of the concentration of structural units. The activity terms aiS and aiP in the equation can be replaced by the mole fraction of structural units, and a linear fitting between the surface tension of slag and the logarithm of structural parameters is established, as shown in Figure 10.
The fitting coefficients of surface tension with the linear equations of lnQ0(Si)+Q1(Si), lnQ2(Si)+Q3(Si), lnNBO/Si, and lnNBO/T are 0.756, 0.715, 0.969, and 0.974, respectively. The comparison shows that the linear correlation between surface tension and lnQ0(Si)+Q1(Si), lnQ2(Si)+Q3(Si) is weak; however, the linear relationship between surface tension and lnNBO/Si and lnNBO/T is strong. This indicates that surface tension has a stronger dependence on the overall microstructure of the slag; that is, lowly polymerized slag results in high surface tensions.
Figure 11 shows the microstructure analysis of the variation of surface tension. When the CaO content is 5%, many complex [SiO4]-tetrahedra in the slag are observed with a larger average ion radius and lower interaction force between ions, resulting in a lower surface tension at this stage. As CaO replaces FeO, an increase in O2− breaks the bridging oxygen bond, causing a gradual depolymerization of complex [SiO4]-tetrahedrons into the simple structural units. The depolymerization of [SiO4]-tetrahedra results in a decrease in the average ion radius of the slag and an increase in the interaction force between ions, thereby causing an increase in surface tension. Furthermore, the increase of O2− leads to an increase in the number of ions in the slag, resulting in a decrease in the distance between ions and an increase in the interaction force. Therefore, there is a significant increase in slag surface tension with increasing CaO content.

5. Conclusions

The viscosity, surface tension, and Raman spectra of CaO-SiO2-FexO slag were investigated to identify the structural evolution and property transformation of the converter slag during the smelting process. The typical conclusions are summarized as follows:
(1) The change in slag structure can be divided into two stages with the substitution of CaO for FeO. When the CaO content in the slag is less than 35%, the complex [SiO4]-tetrahedra depolymerizes into simple structural units. When the CaO content exceeds 35%, most of the complex [SiO4]-tetrahedra are depolymerized, and only a portion of [SiO4]-tetrahedra continue to depolymerize. Meanwhile, Fe3+ ions combine with O2− ions to form more [FeO4]-tetrahedra, resulting in an increase in the [FeO4]/[FeO6] ratio. Nevertheless, the slag is still in a state of depolymerization as a whole.
(2) When the addition of CaO is less than 35%, as CaO replaces FeO, the dissociated O2− in the melt preferentially destroys the Si-O-Si bonds due to its strong affinity with Si4+, resulting in a significant depolymerization of highly polymerized [SiO4]-tetrahedra. The depolymerized slag reduces its activation energy of viscous flow and increases the ion interaction force, resulting in a decrease in slag viscosity and an increase in surface tension.
(3) When CaO content exceeds 35%, O2− not only destroys the Si-O-Si bonds in Q3(Si) and Q2(Si) units but also combines with Fe3+ to form more [FeO4]-tetrahedrons. Consequently, as CaO replaces FeO, the viscosity and surface tension of the slag still maintain the original trend, but the reduction amplitude of viscosity and the increase amplitude of surface tension decrease.

Author Contributions

Conceptualization, R.Z. and Y.M.; methodology, R.Z.; software, C.L.; validation, Z.D., T.-A.Z. and C.L.; formal analysis, R.Z.; investigation, C.L.; resources, T.-A.Z.; data curation, Z.D.; writing—original draft preparation, R.Z.; writing—review and editing, Y.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 52304323), the Open Fund of State Key Laboratory of Vanadium and Titanium Resources Comprehensive Utilization (grand number 2023P4FZG07A), the Fundamental Research Funds for the Central Universities (grant number N232405-06), the Northeast University Postdoctoral Foundation (grant number 20230309), and the Jianlong Group-Northeastern University Youth Science and Technology Innovation Fund (2023012600001).

Data Availability Statement

The data presented in this study are presented in figures and tables in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xua, R.Z.; Zhang, J.L.; Han, W.H.; Chang, Z.Y.; Jiao, K.X. Effect of BaO and Na2O on the viscosity and structure of blast furnace slag. Ironmak. Steelmak. 2020, 47, 168–172. [Google Scholar] [CrossRef]
  2. Geng, X.; Li, B.; Jiang, Z. Effect of TiO2 on viscosity and structure for CaO-SiO2-Al2O3-based mold flux. Calphad 2023, 81, 102545. [Google Scholar] [CrossRef]
  3. Gi, H.K.; Il, S. Effects of alkali metal cations and fluoride anions on viscosity and structure of CaO-SiO2-Al2O3 system. J. Mater. Res. Technol. 2020, 20, 2335–2347. [Google Scholar] [CrossRef]
  4. Wen, Y.; Shu, Q.; Lin, Y.; Fabritius, T. Effect of SiO2 Content and Mass Ratio of CaO to Al2O3 on the viscosity and structure of CaO-Al2O3-B2O3-SiO2 Slags. ISIJ Int. 2023, 63, 1–9. [Google Scholar] [CrossRef]
  5. Gao, Q.; Min, Y.; Jiang, M.F. The temperature and structure dependence of surface tension of CaO–SiO2–Na2O–CaF2 mold fluxes. Metall. Mater. Trans. B 2018, 49, 1302–1310. [Google Scholar] [CrossRef]
  6. Sukenaga, S.; Haruki, S.; Nomoto, Y.; Saito, N.; Nakashima, K. Density and surface tension of CaO–SiO2–Al2O3–R2O (R=Li, Na, K) melts. ISIJ Int. 2011, 51, 1285–1289. [Google Scholar] [CrossRef]
  7. Soledade, M.; Santos, C.S.; Reis, J.C.R. Examination of the Butler equation for the surface tension of liquid mixtures. ACS Omega 2021, 6, 21571–21578. [Google Scholar] [CrossRef]
  8. Wang, Y.Z.; Zhang, Y.; Zhang, H.W. Characteristics of oxygen top blown converter steelmaking method. In The Process and Equipment of Oxygen Top Blown Converter Steelmaking, 2nd ed.; Zhao, P.D., Ed.; Metallurgical Industry Press: Beijing, China, 2001; p. 4. [Google Scholar]
  9. Zhang, R.; Wang, Y.; Zhao, X.; Jia, J.X.; Liu, C.J.; Min, Y. Structure and viscosity of molten CaO-SiO2-FexO slag during the early period of basic oxygen steelmaking. Metall. Mater. Trans. B 2020, 51, 2021–2029. [Google Scholar] [CrossRef]
  10. Santos, R.M.; Ling, D.; Sarvaramini, A.; Guo, M.; Elsen, J.; Larachi, F.; Beaudoin, G.; Blanpain, B.; Gerven, T.V. Stabilization of basic oxygen furnace slag by hot-stage carbonation treatment. Chem. Eng. J. 2012, 203, 239–250. [Google Scholar] [CrossRef]
  11. Yuan, F.; Zhao, Z.; Zhang, Y.; Wu, T. Effect of Al2O3 content on the viscosity and structure of CaO-SiO2-Cr2O3-Al2O3 slags. Int. J. Miner. Metall. Mater 2022, 29, 1522–1531. [Google Scholar] [CrossRef]
  12. Zhang, X.Y.; Li, X.M.; Xing, X.D. Effect of B2O3 content on viscosity and structure of SiO2−MgO−FeO-based slag. Trans. Nonferrous Met. Soc. China 2022, 32, 2403–2413. [Google Scholar] [CrossRef]
  13. Shen, Y.; Chong, J.; Huang, Z.; Tian, J.; Zhang, W.; Tang, X.; Ding, W.; Du, X. Viscosity and structure of a CaO-SiO2-FeO-MgO system during a modified process from nickel slag by CaO. Materials 2019, 12, 2562. [Google Scholar] [CrossRef] [PubMed]
  14. Rüssel, C.; Wiedenroth, A. The effect of glass composition on the thermodynamics of the Fe2+/Fe3+ equilibrium and the iron diffusivity in Na2O/MgO/CaO/Al2O3/SiO2 melts. Chem. Geol. 2004, 213, 125–135. [Google Scholar] [CrossRef]
  15. Wu, Y.Q.; Jiang, G.C.; You, J.L.; Hou, H.Y.; Chen, H. Raman scattering coefficient of symmetrical stretching mode of silicate melt units. Acta Phys. Sin. 2005, 54, 961–966. [Google Scholar] [CrossRef]
  16. Rajavaram, R.; Kim, H.; Park, J.; Lee, C.H.; Lee, J. Bridging between the physical properties: Structure and density of CaO–SiO2–Al2O3 melts at CaO/SiO2=1.3 and different mole% of Al2O3. Ceram. Int. 2019, 45, 19409–19414. [Google Scholar] [CrossRef]
  17. Fang, J.; Pang, Z.; Xing, X.; Xu, R. Thermodynamic properties, viscosity, and structure of CaO-SiO2–MgO-Al2O3-TiO2-based. Materials 2021, 14, 124. [Google Scholar] [CrossRef]
  18. Mysen, B.O.; Virgo, D.; Seifert, F.A. The structure of silicate melts: Implications for chemical and physical properties of natural magma. Rev. Geophys. Space Phys. 1982, 20, 353–383. [Google Scholar] [CrossRef]
  19. Mysen, B.O.; Virgo, D.; Scarfe, C.M. Relations between the anionic structure and viscosity of silicate melts-a Raman spectroscopic study. Am. Mineral. 1980, 65, 690–710. [Google Scholar]
  20. Yang, J.; Zhang, J.Q.; Ostrovski, O.; Zhang, C.; Cai, D.X. Effects of fluorine on solidification, viscosity, structure, and heat transfer of CaO–Al2O3-based mold fluxes. Metall. Mater. Trans. B 2019, 50, 1766–1772. [Google Scholar] [CrossRef]
  21. Qin, J.; Liu, W.; Wu, H.; Wang, J.; Xue, Q.; Zhao, H.; Zuo, H. A comprehensive investigation on the viscosity and structure of CaO-SiO2-Al2O3-MgO-BaO slag with different Al2O3/SiO2 ratios. J. Mol. Liq. 2022, 365, 120060. [Google Scholar] [CrossRef]
  22. Zhou, S. Surface electrostatic force in presence of dimer counter-ion. J. Mol. Liq. 2021, 328, 115225. [Google Scholar] [CrossRef]
  23. Butler, J. The thermodynamics of the surface of solutions. Proc. R. Soc. Lond. A 1932, 135, 348–375. [Google Scholar]
Figure 1. Typical X-ray diffraction pattern of quenched sample.
Figure 1. Typical X-ray diffraction pattern of quenched sample.
Metals 13 01869 g001
Figure 2. Schematic diagram of RTW-10 melt physical property tester.
Figure 2. Schematic diagram of RTW-10 melt physical property tester.
Metals 13 01869 g002
Figure 3. Viscosity–temperature curves of the slags with varying CaO/FexO: (a) different temperatures; (b) linear regression of lnη and T−1.
Figure 3. Viscosity–temperature curves of the slags with varying CaO/FexO: (a) different temperatures; (b) linear regression of lnη and T−1.
Metals 13 01869 g003
Figure 4. Surface tension curves of the slags with varying CaO/FexO.
Figure 4. Surface tension curves of the slags with varying CaO/FexO.
Metals 13 01869 g004
Figure 5. Raman spectra of the CaO-SiO2-FexO slags with different CaO/FexO ratios.
Figure 5. Raman spectra of the CaO-SiO2-FexO slags with different CaO/FexO ratios.
Metals 13 01869 g005
Figure 6. Deconvolution results of the Raman spectra of CaO-SiO2-FexO slags: (a) 5 wt.%CaO; (b) 15 wt.%CaO; (c) 25 wt.%CaO; (d) 35 wt.%CaO; (e) 40 wt.%CaO; (f) 50 wt.%CaO.
Figure 6. Deconvolution results of the Raman spectra of CaO-SiO2-FexO slags: (a) 5 wt.%CaO; (b) 15 wt.%CaO; (c) 25 wt.%CaO; (d) 35 wt.%CaO; (e) 40 wt.%CaO; (f) 50 wt.%CaO.
Metals 13 01869 g006
Figure 7. Relative area fractions of the structural units in the CaO-SiO2-FexO slag.
Figure 7. Relative area fractions of the structural units in the CaO-SiO2-FexO slag.
Metals 13 01869 g007
Figure 8. Linear relationship between the activation energy of viscous flow and the structural parameters: (a) Q0(Si); (b) Q0(Si)+Q1(Si)+Q2(Si); (c) Q3(Si)+Q4(Si); (d) NBO/T.
Figure 8. Linear relationship between the activation energy of viscous flow and the structural parameters: (a) Q0(Si); (b) Q0(Si)+Q1(Si)+Q2(Si); (c) Q3(Si)+Q4(Si); (d) NBO/T.
Metals 13 01869 g008
Figure 9. Microscopic analysis of changes in slag viscosity.
Figure 9. Microscopic analysis of changes in slag viscosity.
Metals 13 01869 g009
Figure 10. Linear relationship between the surface tension and the structural parameters: (a) Q0(Si)+Q1(Si); (b) Q2(Si)+Q3(Si); (c) NBO/TSi; (d) NBO/T.
Figure 10. Linear relationship between the surface tension and the structural parameters: (a) Q0(Si)+Q1(Si); (b) Q2(Si)+Q3(Si); (c) NBO/TSi; (d) NBO/T.
Metals 13 01869 g010
Figure 11. Microscopic analysis of changes in surface tension of slag.
Figure 11. Microscopic analysis of changes in surface tension of slag.
Metals 13 01869 g011
Table 1. Composition of CaO-SiO2-FeO-Fe2O3 slags (mass fraction, %).
Table 1. Composition of CaO-SiO2-FeO-Fe2O3 slags (mass fraction, %).
No.CaOSiO2FeOFe2O3Fe2+/T.Fe
CSF16.4738.4133.6621.460.63
CSF214.3440.6630.0314.970.69
CSF324.3741.5523.0910.990.70
CSF436.4139.6614.579.360.63
CSF541.5640.8311.995.620.70
CSF652.1437.655.634.580.58
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

Zhang, R.; Min, Y.; Liu, C.; Zhang, T.-A.; Dou, Z. The Effect of Substituting FeO with CaO on the Rheological and Surface Properties of Silicate Melts. Metals 2023, 13, 1869. https://doi.org/10.3390/met13111869

AMA Style

Zhang R, Min Y, Liu C, Zhang T-A, Dou Z. The Effect of Substituting FeO with CaO on the Rheological and Surface Properties of Silicate Melts. Metals. 2023; 13(11):1869. https://doi.org/10.3390/met13111869

Chicago/Turabian Style

Zhang, Rui, Yi Min, Chengjun Liu, Ting-An Zhang, and Zhihe Dou. 2023. "The Effect of Substituting FeO with CaO on the Rheological and Surface Properties of Silicate Melts" Metals 13, no. 11: 1869. https://doi.org/10.3390/met13111869

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