Next Article in Journal
Properties of Coatings Based on Calcium Phosphate and Their Effect on Cytocompatibility and Bioactivity of Titanium Nickelide
Next Article in Special Issue
Microstructure and Mechanical Properties of Cast Al-Si-Cu-Mg-Ni-Cr Alloys: Effects of Time and Temperature on Two-Stage Solution Treatment and Ageing
Previous Article in Journal
Impact of Irradiation on the Adhesive Performance of Resin-Based Dental Biomaterials: A Systematic Review of Laboratory Studies
Previous Article in Special Issue
Effect of Solution Temperature on Microstructure and Properties of Thixotropic Back-Extruded Tin–Bronze Shaft Sleeve
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of TiO2, Al2O3, and Basicity on Viscosity and Structure of High Titanium-Bearing Blast Furnace Slag

1
School of Materials and Metallurgy, University of Science and Technology Liaoning, Anshan 114051, China
2
Fujian Sansteel (Group) Co., Ltd., Sanming 365000, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(7), 2575; https://doi.org/10.3390/ma16072575
Submission received: 20 February 2023 / Revised: 13 March 2023 / Accepted: 21 March 2023 / Published: 24 March 2023
(This article belongs to the Special Issue Semisolid Processing and Squeeze Casting of Alloys and Composites)

Abstract

:
The viscosity of high-titanium blast furnace slag with different TiO2 content, Al2O3 content, and basicity was measured at 1653–1773 K using the rotational cylinder method. The phase composition of the slag is measured by XRD. Phase diagram of the slags is calculated by FactSage software. Ionic network structure of the slags is analyzed by FT–IR. Results show that TiO2 depolymerizes the silicate network structure, reducing viscosity at high temperature, while increasing Al2O3 content generates a more complicated silicate, increasing viscosity. Basicity affects viscosity, with higher basicity resulting in lower viscosity above 1733 K. Perovskite significantly affects the viscosity of slag. This study provides an in-depth understanding of the relationship between the composition and viscosity of high-titanium blast furnace slag, which is very important for improving production efficiency.

1. Introduction

Viscosity is an important physical property of slag to determine the stability and productivity of the blast furnace iron-making. Conventional blast furnace slag mainly contains CaO, SiO2, MgO, and Al2O3. When vanadium–titanium ore is used as a raw material for blast furnace iron-making, the softening and melting zone of the furnace changed [1,2,3,4,5,6]. Most of the TiO2 enters the blast furnace slag, making it sticky and causing difficulties in separating slag and iron in the furnace hearth. The viscosity and structure of the blast furnace slag changed due to the addition of TiO2 [7,8,9].
High titanium-bearing blast furnace slag is the main type of blast furnace slag produced by vanadium and titanium ore in the Panzhihua area during blast furnace smelting, which is mainly composed of oxides such as CaO, SiO2, MgO, Al2O3, and TiO2, and when the temperature is constant, the viscosity of the slag mainly depends on these oxide components [10,11]. Some investigations on the viscosity of titanium-bearing slag have been carried out. It is generally believed that the viscosity decreases with the increase in TiO2 content [12,13]. However, there are still different opinions about the mechanism of the influence of TiO2. One opinion is that TiO2 as the network former forms the simpler structure of TiO44− monomers. So, the strength of the network structure is weakened [11,14,15]. Another opinion is that TiO2 acting as a basic oxide in the slag depolymerizes the silicate network structure [16,17].
Suitable basicity plays a crucial role in the blast furnace smelting process. Excessive basicity leads to an increase in high-melting-point compounds in the slag, resulting in poor fluidity of blast furnace slag and difficulty in separating slag and iron. If the basicity is too low, the desulfurization capacity of blast furnace slag decreases, and at the same time, the blast furnace slag erodes the furnace lining. On the other hand, it is well-known that Al2O3 is a typical amphoteric oxide. In the CaO-MgO-SiO2-Al2O3 slag system, whether Al2O3 acts as an acidic oxide or as an alkaline oxide depends on the slag composition. In some slags, Al2O3 is more inclined to act as a network former, and the viscosity increases with Al2O3 addition [18,19]. So, the effect of Al2O3 on the viscosity of the slag with different composition is different. In other slags, the viscosity first increases and then decreases with Al2O3 addition [20,21]. In the smelting of vanadium–titanium ore in a blast furnace, the physical and chemical properties, as well as the composition of the liquid phase line temperature and viscosity of the blast furnace slag, constantly change during the smelting process. These parameters have a significant impact on the smooth flow of the smelting process. However, there are a few reports on the physical and chemical properties of high-titanium slag. Therefore, obtaining these physical and chemical parameters through systematic testing is of great practical significance for optimizing the smelting process, reducing energy consumption, and improving production efficiency.
This paper investigates the impact of TiO2, Al2O3, and basicity (R = m(CaO)/m(SiO2)) on the viscosity of CaO-SiO2-MgO-Al2O3-TiO2 slags. Fourier transform infrared (FT–IR) spectra of the water-quenched slag is used to study the relationship between the network structure and viscosity. X-ray diffraction (XRD) analysis is conducted on the water-quenched slag at various temperatures to determine the liquidus isotherm using Factsage 7.2. This information is used to examine the effect of solid precipitates in the slag on viscosity. The study aims to provide a theoretical basis for utilizing vanadium–titanium ore in blast furnace and analyze the impact of these variables on blast furnace slag viscosity.

2. Materials and Methods

Preparing slag by mixing pure oxides (MgO, SiO2, Al2O3, TiO2) and pure CaCO3 as the source of CaO. First, 200 g of slag powder was pre-melted at 1773 K for 1 h, then the sample was removed and cooled to room temperature by water cooling to obtain a water-quenched slag as an experimental sample. About 20 g slag samples were used to carry out the qualitative FT–IR and XRD characterization of the slags and an additional slag sample was used to measure the viscosity. The calculated chemical composition of the slag is shown in Table 1. The design composition of the slag is based on the actual composition of the Pangao steel plant [8,9,10].
In this study, the slag viscosity was measured using the rotating-cylinder method [22]. The experimental equipment used is a high temperature melt property tester (RT-3, Mingjian Hi-Tech Industrial Co., Suzhou, China). Figure 1a shows the experimental apparatus, which consists of an electric resistance furnace equipped with U-shaped MoSi2 heating elements for system heating. A Mo crucible with a height of 80 mm and an inner diameter of 40 mm was used to hold a 140 g sample, which was heated at a rate of 5 K/min up to 1773 K with a constant flow of Ar gas (500 mL/min). Viscosity measurement was conducted at every 40 K interval during cooling, with an equilibration time of 30 min at each temperature. The rotating spindle, as detailed in Figure 1b, was set to rotate at a speed of 300 rev/min, and five viscosity measurements were taken at the same temperature. The viscosity reported in this paper is an average of these five measurements.

3. Results and Discussion

The viscosity of the CaO-SiO2-MgO-Al2O3-TiO2 slags with different basicity, TiO2 content, and Al2O3 content are shown in the Table 2. The viscosity increases with the decrease in temperature. When the temperature decreases to 1653 K, the viscosity shows a large increasing trend in some slag systems.

3.1. Effect of TiO2 on Viscosity and Structure

Figure 2 shows the effect of TiO2 content on viscosity of the CaO-SiO2-8 mass% MgO-14 mass% Al2O3-TiO2 slag with R = 1.10. The viscosity decreases with the increase in TiO2 content from 10 mass% to 30 mass% when the temperature is greater than 1693 K, which is consistent with previous study [11,15]. However, at 1653 K, the viscosity first increases and then decreases with the increase in TiO2 content, and it shows a maximum when TiO2 content is 20 mass%. This is different from the previous studies, where it was found that the viscosity of CaO-SiO2-TiO2 slag is the highest when the TiO2 content is 25% by mass [15].
Figure 3 shows the isotherm phase diagram calculated by FactSage [23]. At temperatures above 1693 K, the composition of the slag is entirely in liquid phase, and its viscosity is controlled by the network structure of the liquid phase. As the TiO2 content in the slag increases, the viscosity exhibits a decreasing trend. This can be attributed to the fact that the ionic radius of Ti4+ is about 1.5 times larger than that of Si4+, resulting in a weaker bond between Ti4+ and O2− compared to that between Si4+ and O2− [8]. Consequently, the formation of TiO44− monomers weakens the network structure [15]. On the other hand, the effect of Al2O3 content on viscosity is different from that of TiO2, despite the larger ionic radius of Al3+ compared to that of Si4+ in the slag [18,24]. Therefore, the reason for TiO2 addition leading to a decrease in viscosity is not yet clear, although it is possible that TiO2 may exist as the basic oxide and depolymerize the network structure in the present slag systems.
It can be seen from Figure 3 that liquidus temperature of the slags is greater than 1653 K. Perovskite is present in the slag at 1653 K, according to the XRD curves of the slags in Figure 4. So, the viscosity should be affected by the solid phase precipitated from the slag at low temperatures. As can be seen from Figure 3, the 1653 K isothermal curve is much further away from the composition points of slags containing 15 mass% to 25 mass% TiO2 than from slags containing 10 mass% and 30 mass%. It can be considered that the former have more solid phase in the slag than the later. The content of the solid phase plays a key role in the slag viscosity at 1653 K. Hence, the viscosity increases first and then decreases with TiO2 content from 10 mass% to 30 mass%.

3.2. Effect of Basicity on Viscosity and Structure

Figure 5 illustrates how basicity affects the viscosity of the CaO-SiO2-Al2O3-MgO-25 mass% TiO2 slag. As shown, the viscosity of the slag decreases with increasing basicity when the temperature is greater than 1733 K. This can be attributed to the network structure, as indicated in Figure 3, where the liquidus temperature of the slag with a basicity ranging from 1.00 to 1.20 is lower than 1733 K. CaO serves as a network modifier, which depolymerizes the network structure, whereas SiO2 acts as a network former, which polymerizes the network structure. The rise in basicity corresponds to an increase in CaO content and a decrease in SiO2 content. In other words, an increase in basicity reduces both the size and number of the silicate network structure, leading to a decrease in viscosity.
Figure 6 shows the XRD results of the CaO-SiO2-16 mass% Al2O3-10 mass% MgO-25 mass% TiO2 slag system at different basicity levels at 1653 K. The solid phase of the slag contains calcium titanate and spinel phases. As the basicity increases from 1.00 to 1.20, the content of the solid phase in the slag increases significantly, leading to an increase in viscosity at 1653 K. In particular, although there are solid phase points with an basicity of 1.00 in the slag at 1653 K, the viscosity is still low. This may be because the solid phase is mainly dispersed in the liquid phase, and when the amount of solid phase is small, it does not form large aggregates. The formation of large aggregates by the solid phase has almost no effect on the viscosity.

3.3. Effect of Al2O3 on Viscosity and Structure

Figure 7 shows the effect of Al2O3 on viscosity of the CaO-SiO2-Al2O3-MgO-25 mass% TiO2 slag with R = 1.10. The viscosity increases with the increase of Al2O3 content in the slag at 1733 and 1773 K, and the viscosity decreases at 1653 K. At 1693 K when the Al2O3 content increases, the viscosity of the slag with 8 mass% MgO decreases and the viscosity of the slag with 10 mass% MgO increases, with the viscosity of the two contents changing in opposite patterns. It is widely believed that Al2O3 acts as an amphoteric oxide in the slag [25,26,27,28]. The amphoteric oxide may show different properties in different slag system [21,26]. In the present study, Al2O3 may be expected, as a network former, to cause an increase in the viscosity at high temperature.
The viscosity of the sample undergoes a significant increase when the temperature drops to 1653 K, which can be attributed to the solid phase appearing, as indicated in Figure 8 calculated by FactSage [23]. As illustrated in Figure 8a, at 1653 K, the liquidus temperature decreases with the increase in Al2O3 content, when the MgO content is 8 mass%. The reduction in solid phase content may be the primary reason for the viscosity decrease with the addition of Al2O3 content. When the Al2O3 content increases from 14 mass% to 18 mass% at 10 mass% MgO content, the primary crystal zone undergoes a change, as observed in Figure 8b. The XRD results depicted in Figure 9 show that the solid phase is converted from perovskite to perovskite and spinel when the Al2O3 content is added. In some instances, the intensity of the perovskite peak reduces with the introduction of the spinel. When the content of spinel in the slag is low, the effect of perovskite on viscosity is potentially greater than that of spinel. Hence, at 1653 K and 10 mass% MgO content, the addition of Al2O3 content induces a decrease in viscosity.

3.4. FT–IR Spectra of the Slag

In order to further understand the influence mechanism of the slag compositions on the viscosity, FT–IR spectra of the quenched slag at 1773 K were obtained. Figure 10 shows the FT–IR spectra of the slag containing different TiO2 content. The FT–IR spectrum of the slag can be divided into three crucial ranges according to the wave number. These ranges represent [SiO4]-tetrahedra between 1200 cm−1 and 800 cm−1, [AlO4]-tetrahedra between 730 cm−1 and 630 cm−1, and [Al-O-Si]-rocking between 480 cm−1 and 410 cm−1, respectively [29,30].
The depth of the FT–IR spectra between 1200 cm−1 and 800 cm−1 decreases with the increase in TiO2 content, which indicates that the network structure of [SiO4]-tetrahedra becomes simpler with TiO2 content addition. The depth of the peak representing the [Al-O-Si]-rocking is also decreased with the increase in TiO2 content. The effect of TiO2 content on [AlO4]-tetrahedra is insignificant. According to previous study, the existent forms of TiO2 in the slag include two types [12,31]. TiO2 as a basic oxide provides Ti4+ to depolymerize network structure or it as an acidic oxide forms simple network structure of TiO44−. In the present study, TiO2 might be more inclined to existing as a network modifier and depolymerizes the silicate network structure. So, the viscosity decreases with TiO2 addition.
Figure 11 shows the FT–IR spectra of the slag with different basicity. The depth of the FT–IR spectra between 1200 cm−1 and 800 cm−1 decreases with the increase in basicity. Sometime, the depth of FT–IR spectra representing the [AlO4]-tetrahedra has hardly changed. It is similar to the effect of TiO2 on the network structure. The network structure of [SiO4]-tetrahedra should become simpler with the increase in basicity, which is consistent with the description in Figure 5.
The FT–IR spectra of the slag with different Al2O3 contents are shown in Figure 12. The depth of the FT–IR spectra characterizing the [SiO4]-tetrahedra has hardly changed and it characterizing [AlO4]-tetrahedra significantly increases with the Al2O3 content from 14 mass% to 16 mass%. As Al2O3 content continues to 18 mass%, the depth of the FT–IR spectra characterizing the [SiO4]-tetrahedra does not only increase but also the central position of the peak of the FT–IR spectra has changed. Sometimes, the depth of the FT–IR spectra characterizing [AlO4]-tetrahedra decreases.
It is easily understandable that the network structure of [AlO4]-tetrahedra in the slag system increases as the Al2O3 content serves as an acidic oxide. However, when the Al2O3 content reaches 18 mass%, the network structure of [AlO4]-tetrahedra is reduced and the network structure of [SiO4]-tetrahedra becomes more complicated. This is because some of the Si4+ in [SiO4]-tetrahedra are replaced by Al3+, and Al3+ still exist in the form of [SiO4]-tetrahedra. As a result, a more intricate silicate network structure is formed while the network structure of [AlO4]-tetrahedra is reduced. It is evident that the effect of the silicate network structure on viscosity should be more pronounced than that of [AlO4]-tetrahedra. According to Figure 2 and Figure 12, the increase in viscosity with Al2O3 content from 14 mass% to 16 mass% is attributed to an increased proportion of [AlO4]-tetrahedra in the network structure, whereas the increase in viscosity with Al2O3 content from 16 mass% to 18 mass% is attributed to the more complex silicate network structure generated.

4. Conclusions

In this study, the viscosity of CaO-SiO2-MgO-Al2O3-TiO2 slag with different TiO2 content (10–30 mass%), Al2O3 content (14–16 mass%), and basicity (1.00–1.20) was measured. FactSage was used to draw the slag phase diagram and X-ray diffraction analysis was performed to understand the effect of calcium titanate and spinel solid phases on slag viscosity, and the relationship between network structure and viscosity was studied through FT–IR spectra.
(1)
TiO2 might be more inclined to existing as a network modifier, and depolymerizes the silicate network structure in liquid slag. So, the viscosity decreases with the TiO2 content from 10 mass% to 30 mass% at high temperature.
(2)
The viscosity of the slag increases as the content of Al2O3 increases from 14 mass% to 16 mass%. This is due to the formation of a network structure of [AlO4]-tetrahedra. However, when the Al2O3 content increases from 16 mass % to 18 mass %, there is a partial replacement of Si4+ in [SiO4]-tetrahedra with Al3+, while some Al3+ still exists in the form of [SiO4]-tetrahedra. As a result, the increase in viscosity can be attributed to the generation of a more complex silicate structure.
(3)
The basicity of slag affects its viscosity, with higher basicity resulting in lower viscosity at temperatures above 1733 K. The decrease in viscosity is due to the reduction in size and number of silicate network structure caused by an increase in CaO content and a decrease in SiO2 content. The increase in solid phase content with increasing basicity results in an increase in viscosity at 1653 K, but when the amount of solid phase is small, it has almost no effect on the viscosity.
(4)
At 1653 K, the solid phase in the current slag system includes perovskite and spinel, with perovskite playing an important role in viscosity.

Author Contributions

W.Z., D.L. and C.S. contributed to the material synthesis, performed the experiments, material characterization, data analysis, and paper writing; T.L. and S.Y. contributed to the data analysis, paper writing, and design of the experiment. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Programs of the National Natural Science Foundation of China (Nos. 52174319 and 51474125).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, S.; Tang, W.; Xue, X. Effect of TiO2 on the sintering behavior of low-grade vanadiferous titanomagnetite ore. Materials 2021, 14, 4376. [Google Scholar] [CrossRef] [PubMed]
  2. Li, J.; Men, J.; Yang, S.; Zhou, M. Influence of fuel level on properties, productivity, and mineralogy of russian vanadiferous titanomagnetite sinter. Materials 2021, 14, 6258. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, S.; Zhou, M.; Jiang, T.; Wang, Y.; Xue, X. Effect of basicity on sintering behavior of low-titanium vanadium-titanium magnetite. Trans. Nonferrous Met. Soc. China 2015, 25, 2087–2094. [Google Scholar] [CrossRef]
  4. Yang, S.; Zhou, M.; Xue, X.; Jiang, T.; Sun, C. Isothermal reduction kinetics of chromium-bearing vanadium–titanium sinter reduced with CO gas at 1173 K. JOM 2019, 71, 2812–2820. [Google Scholar] [CrossRef]
  5. Yang, S.; Gao, Y.; Xue, X.; Li, B. Influence of titanium on transformation behaviour during continuous cooling of boron microalloyed steels. Ironmak. Steelmak. 2018, 45, 959–968. [Google Scholar] [CrossRef]
  6. Yang, S.; Zhou, M.; Jiang, T.; Xue, X. Isothermal reduction kinetics and mineral phase of chromium-bearing vanadium–titanium sinter reduced with CO gas at 873–1273 K. Int. J. Miner. Metall. Mater. 2018, 25, 145–152. [Google Scholar] [CrossRef]
  7. Shankar, A.; Görnerup, M.; Lahiri, A.K.; Seetharaman, S. Experimental Investigation of the viscosities in CaO-SiO2-MgO-Al2O3 and CaO-SiO2-MgO-Al2O3-TiO2 slags. Metall. Mater. Trans. B 2007, 38, 911–915. [Google Scholar] [CrossRef]
  8. Dong, X.J.; Sun, H.Y.; She, X.F.; Xue, Q.G.; Wang, J.S. Viscosity and viscosity estimation model of fully liquid slags in TiO2-Al2O3-CaO-SiO2 and TiO2-Al2O3-CaO-SiO2-MgO systems with high TiO2 concentration and low mass ratio of CaO to SiO2. Ironmak. Steelmak. 2014, 41, 99–106. [Google Scholar] [CrossRef]
  9. Shi, J.; Sun, L.; Qiu, J.; Wang, Z.; Zhang, B.; Jiang, M. Experimental determination of the phase diagram for CaO-SiO2-MgO-10%Al2O3-5%TiO2 system. ISIJ Int. 2016, 56, 1124–1131. [Google Scholar] [CrossRef] [Green Version]
  10. Yan, Z.; Lv, X.; He, W.; Xu, J. Effect of TiO2 on the liquid zone and apparent viscosity of SiO2-CaO-8wt%mgo-14wt%Al2O3 system. ISIJ Int. 2016, 57, 31–36. [Google Scholar] [CrossRef] [Green Version]
  11. Yan, Z.; Lv, X.; Zhang, J.; Qin, Y.; Bai, C. Influence of mgo, Al2O3 and CaO/SiO2 on the viscosity of blast furnace type slag with high Al2O3 and 5 wt-% TiO2. Can. Metall. Q. 2016, 55, 186–194. [Google Scholar] [CrossRef]
  12. Jong, B.K.; Sohn, I. Effect of SiO2/Al2O3 and TiO2/SiO2 Ratios on the viscosity and structure of the TiO2-MnO-SiO2-Al2O3 welding flux system. ISIJ Int. 2014, 54, 2050–2058. [Google Scholar]
  13. Zhang, S.F.; Zhang, X.; Liu, W.; Lv, X.; Bai, C.G.; Wang, L. Relationship between structure and viscosity of CaO-SiO2-Al2O3-MgO-TiO2 slag. J. Non-Cryst. Solids 2014, 402, 214–222. [Google Scholar] [CrossRef]
  14. Zhen, Y.L.; Zhang, G.H.; Chou, K.C. Influence of Al2O3/TiO2 ratio on viscosities and structure of CaO-MgO-Al2O3-SiO2-TiO2 melts. ISIJ Int. 2014, 54, 985–989. [Google Scholar] [CrossRef] [Green Version]
  15. Zheng, K.; Zhang, Z.; Liu, L.; Wang, X. Investigation of the viscosity and structural properties of CaO-SiO2-TiO2 slags. Metall. Mater. Trans. B 2014, 45, 1389–1397. [Google Scholar] [CrossRef]
  16. Liao, J.L.; Li, J.; Wang, X.D.; Zhang, Z.T. Influence of TiO2 and basicity on viscosity of Ti bearing slag. Ironmak. Steelmak. 2012, 39, 133–139. [Google Scholar] [CrossRef]
  17. Sohn, I.; Wang, W.L.; Matsuura, H.; Tsukihashi, F.; Min, D.J. Influence of TiO2 on the viscous behavior of calcium silicate melts containing 17 mass% Al2O3 and 10 mass% MgO. ISIJ Int. 2012, 52, 158–160. [Google Scholar] [CrossRef] [Green Version]
  18. Shiau, J.S.; Liu, S.H.; Ho, C.K. Effect of magnesium and aluminum oxides on fluidity of final blast furnace slag and its application. Mater. Trans. 2012, 53, 1449–1455. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, X.; Tao, J.; Xue, X.; Hu, B. Influence of MgO/Al2O3 ratio on viscosity of blast furnace slag with high Al2O3 content. Steel Res. Int. 2016, 87, 87–94. [Google Scholar] [CrossRef]
  20. Yan, Z.; Lv, X.; Liang, D.; Zhang, J.; Bai, C. Transition of blast furnace slag from silicates-based to aluminates-based: Viscosity. Metall. Mater. Trans. B 2017, 48, 1092–1099. [Google Scholar] [CrossRef]
  21. Park, J.H.; Min, D.J.; Song, H.S. Amphoteric behavior of alumina in viscous flow and structure of CaO-SiO2 (-MgO)-Al2O3 slags. Metall. Mater. Trans. B 2004, 35, 269–275. [Google Scholar] [CrossRef]
  22. Sun, C.Y.; Liu, X.H.; Jing, L.I.; Yin, X.T.; Song, S.; Wang, Q. Influence of Al2O3 and mgo on the viscosity and stability of CaO-MgO-SiO2-Al2O3 slags with cao/sio_2=1.0. ISIJ Int. 2017, 57, 978–982. [Google Scholar] [CrossRef] [Green Version]
  23. Bale, C.W.; Pelton, A.D.; Thompson, W.T.; Eriksson, G. FactSage, Ecole Polytechnique, Montreal. 2001. Available online: http://www.crct.polymtl.ca (accessed on 10 November 2022).
  24. Hu, X.J.; Ren, Z.S.; Zhang, G.H.; Wang, L.J.; Chou, A.K.C. A model for estimating the viscosity of blast furnace slags with optical basicity. Int. J. Miner. Metall. Mater. 2012, 19, 1088–1092. [Google Scholar] [CrossRef]
  25. Masashi, N.; Toshihiro, T.; Joonho, L.; Tateo, U. Evaluation of viscosity of molten SiO2-CaO-MgO-Al2O3 slags in blast furnace operation. ISIJ Int. 2004, 44, 2115–2119. [Google Scholar]
  26. Kim, H.; Matsuura, H.; Tsukihashi, F.; Wang, W.; Min, D.J.; Sohn, I. Effect of Al2O3 and CaO/SiO2 on the viscosity of calcium-silicate-based Slags containing 10 mass Pct MgO. Metall. Mater. Trans. B 2012, 44, 5–12. [Google Scholar] [CrossRef]
  27. Park, J.; Kim, H.; Min, D. Novel Approach to link between viscosity and structure of silicate melts via darken’s excess stability function: Focus on the amphoteric behavior of alumina. Metall. Mater. Trans. B 2008, 39, 150–153. [Google Scholar] [CrossRef]
  28. Sunahara, K.; Nakano, K.; Hoshi, M.; Inada, T.; Komatsu, S.; Yamamoto, T. Effect of high Al2O3 slag on the blast furnace operations. ISIJ Int. 2008, 48, 420–429. [Google Scholar] [CrossRef] [Green Version]
  29. Jiao, K.X.; Zhang, J.L.; Wang, Z.Y.; Chen, C.L.; Liu, Y.X. Effect of TiO2 and FeO on the viscosity and structure of blast furnace primary slags. Steel Res. Int. 2016, 88, 1600296. [Google Scholar] [CrossRef]
  30. Sun, Y.; Zheng, K.; Liao, J.; Wang, X.; Zhang, Z. Effect of P2O5 addition on the viscosity and structure of titanium bearing blast furnace slags. ISIJ Int. 2014, 54, 1491–1497. [Google Scholar] [CrossRef] [Green Version]
  31. Park, H.; Park, J.Y.; Kim, G.H.; Sohn, I. Effect of TiO2 on the viscosity and slag structure in blast furnace type slags. Steel Res. Int. 2012, 83, 150–156. [Google Scholar] [CrossRef]
Figure 1. The schematic diagram of viscosity measurements. (a) Experiment apparatus; (b) Mo spindle.
Figure 1. The schematic diagram of viscosity measurements. (a) Experiment apparatus; (b) Mo spindle.
Materials 16 02575 g001
Figure 2. Viscosity of the CaO-SiO2-8 mass% MgO-14 mass% Al2O3-TiO2 (R = 1.10) slag as a function of mass% TiO2.
Figure 2. Viscosity of the CaO-SiO2-8 mass% MgO-14 mass% Al2O3-TiO2 (R = 1.10) slag as a function of mass% TiO2.
Materials 16 02575 g002
Figure 3. Phase diagram of CaO-SiO2-8 mass% MgO-14 mass % Al2O3-TiO2 slag, K.
Figure 3. Phase diagram of CaO-SiO2-8 mass% MgO-14 mass % Al2O3-TiO2 slag, K.
Materials 16 02575 g003
Figure 4. XRD curves of the slag containing different TiO2 content at 1653 K.
Figure 4. XRD curves of the slag containing different TiO2 content at 1653 K.
Materials 16 02575 g004
Figure 5. Viscosity of the CaO-SiO2-MgO-Al2O3-25 mass% TiO2 slag as a function of basicity.
Figure 5. Viscosity of the CaO-SiO2-MgO-Al2O3-25 mass% TiO2 slag as a function of basicity.
Materials 16 02575 g005
Figure 6. XRD curves of the slag containing different basicity at 1653 K.
Figure 6. XRD curves of the slag containing different basicity at 1653 K.
Materials 16 02575 g006
Figure 7. Viscosity of the CaO-SiO2-MgO-Al2O3-25 mass% TiO2 (R = 1.10) slag as a function of mass% Al2O3.
Figure 7. Viscosity of the CaO-SiO2-MgO-Al2O3-25 mass% TiO2 (R = 1.10) slag as a function of mass% Al2O3.
Materials 16 02575 g007
Figure 8. Phase diagram of CaO-SiO2-MgO-Al2O3-25 mass% TiO2 slag, K. (a) 8 mass% MgO; (b) 10 mass% MgO.
Figure 8. Phase diagram of CaO-SiO2-MgO-Al2O3-25 mass% TiO2 slag, K. (a) 8 mass% MgO; (b) 10 mass% MgO.
Materials 16 02575 g008
Figure 9. XRD curves of the slag containing different Al2O3 content at 1653 K.
Figure 9. XRD curves of the slag containing different Al2O3 content at 1653 K.
Materials 16 02575 g009
Figure 10. FT–IR transmittance spectra of the slag containing different TiO2 content at 1773 K.
Figure 10. FT–IR transmittance spectra of the slag containing different TiO2 content at 1773 K.
Materials 16 02575 g010
Figure 11. FT–IR transmittance spectra of the slag containing different basicity at 1773 K.
Figure 11. FT–IR transmittance spectra of the slag containing different basicity at 1773 K.
Materials 16 02575 g011
Figure 12. FT–IR transmittance spectra of the slag containing different Al2O3 content at 1773 K.
Figure 12. FT–IR transmittance spectra of the slag containing different Al2O3 content at 1773 K.
Materials 16 02575 g012
Table 1. The chemical composition of slags.
Table 1. The chemical composition of slags.
Sample
Number
Chemical Compositions (Mass %)R = m(CaO)/m(SiO2)
CaOSiO2MgOAl2O3TiO2
135.6232.388.0014.0010.001.10
233.0030.008.0014.0015.001.10
330.3827.628.0014.0020.001.10
427.7625.248.0014.0025.001.10
525.1422.868.0014.0030.001.10
626.5026.508.0014.0025.001.00
727.1525.858.0014.0025.001.05
828.3524.658.0014.0025.001.15
928.9124.098.0014.0025.001.20
1024.5024.5010.0016.0025.001.00
1125.1023.9010.0016.0025.001.05
1225.6723.3310.0016.0025.001.10
1326.2122.7910.0016.0025.001.15
1426.7322.2710.0016.0025.001.20
1526.7124.298.0016.0025.001.10
1625.6723.338.0018.0025.001.10
1726.7124.2910.0014.0025.001.10
1824.6222.3810.0018.0025.001.10
Table 2. Measured viscosity values of the CaO-SiO2-MgO-Al2O3-TiO2 slags.
Table 2. Measured viscosity values of the CaO-SiO2-MgO-Al2O3-TiO2 slags.
Sample
Number
Viscosity (Pa·s)
1653 K1693 K1733 K1773 K
10.570.430.330.26
20.850.400.300.25
31.470.310.250.22
41.240.270.180.15
50.690.220.170.15
60.280.220.180.16
70.860.220.180.16
81.430.340.160.14
91.570.520.150.13
100.240.210.200.19
110.250.210.190.19
121.180.250.210.18
131.420.180.160.15
141.800.290.150.14
150.930.270.230.20
160.920.290.230.21
171.390.280.190.17
181.130.240.210.19
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

Zhou, W.; Li, T.; Lan, D.; Sun, C.; Yang, S. Influence of TiO2, Al2O3, and Basicity on Viscosity and Structure of High Titanium-Bearing Blast Furnace Slag. Materials 2023, 16, 2575. https://doi.org/10.3390/ma16072575

AMA Style

Zhou W, Li T, Lan D, Sun C, Yang S. Influence of TiO2, Al2O3, and Basicity on Viscosity and Structure of High Titanium-Bearing Blast Furnace Slag. Materials. 2023; 16(7):2575. https://doi.org/10.3390/ma16072575

Chicago/Turabian Style

Zhou, Wenbo, Tingle Li, Dong Lan, Changyu Sun, and Songtao Yang. 2023. "Influence of TiO2, Al2O3, and Basicity on Viscosity and Structure of High Titanium-Bearing Blast Furnace Slag" Materials 16, no. 7: 2575. https://doi.org/10.3390/ma16072575

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