Next Article in Journal
Assessment of Mechanical, Chemical, and Biological Properties of Ti-Nb-Zr Alloy for Medical Applications
Next Article in Special Issue
Effect of Binders on the Crushing Strength of Ferro-Coke
Previous Article in Journal
Spatial Decomposition of a Broadband Pulse Caused by Strong Frequency Dispersion of Sound in Acoustic Metamaterial Superlattice
Previous Article in Special Issue
A Novel Approach for Measuring the Thickness of Refractory of Metallurgical Vessels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamic Properties, Viscosity, and Structure of CaO–SiO2–MgO–Al2O3–TiO2–Based Slag

1
School of Metallurgical Engineering, Xi’an University of Architecture and Technology, Xi’an 710055, China
2
Shannxi Steel Group Hanzhong Iron and Steel Co., Ltd., Hanzhong 724200, China
3
School of Metallurgical and Ecological Engineering, University of Science and Technology Beijing, Beijing 100083, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(1), 124; https://doi.org/10.3390/ma14010124
Submission received: 14 December 2020 / Revised: 24 December 2020 / Accepted: 26 December 2020 / Published: 30 December 2020

Abstract

:
The effect of TiO2 and the MgO/Al2O3 ratio on the viscosity, heat capacity, and enthalpy change of CaO–SiO2–Al2O3–MgO–TiO2 slag at constant heat input was studied. The variation of slag structure was analyzed by the calculation of activation energy and FTIR spectrum measurements. The results showed that the heat capacity and enthalpy change of the slag decreased with the increase of TiO2 content. Under constant heat supply, the fluctuations in slag temperature were relatively apparent, and the temperature of slag increased as the TiO2 content increased. The viscosity of slag decreased due to the increase in slag temperature. Increasing the MgO/Al2O3 ratio could decrease the temperature and viscosity of slag. The effect of increasing the MgO/Al2O3 ratio on the viscosity was more pronounced than the decreasing temperature caused by increasing the MgO/Al2O3 ratio. The apparent activation energy decreased with increasing TiO2 content and MgO/Al2O3 ratio. The Ti–O bonds formed with TiO2 addition, and the Ti–O bonds were weaker than Si–O bonds, which resulted in the decrease in heat capacity and viscosity of slag.

1. Introduction

The viscosity of slag is a vital factor in blast furnace (BF) smelting and has a pivotal role in gas permeability, BF smooth operation, and reduction of iron oxide [1,2,3,4]. The viscosity of slag is affected by its composition. The reserves of vanadium-titanium magnetite (VTM) ores are relatively abundant in China and Australia, and the VTM ores usually contain significant TiO2 [5]. These ores have been used in BF smelting, and as a result the BF slag contains an amount of TiO2. In addition, many steel enterprises adjust the MgO content to improve the fluidity of slag since the high aluminum content of slag increases its viscosity. Both aspects have a great impact on the viscosity and thermal stability of slag. Therefore, it is of great significance to study the effect of TiO2 and the MgO/Al2O3 ratio on the viscosity of BF slag.
Research related to the effect of the slag component on the viscosity have been broadly conducted. Sohn et al. [6] studied the effect of TiO2 from 0 to 10 wt.% on the viscosity of calcium silicate molten slag containing 17 wt.% Al2O3, where TiO2 depolymerized the silicate network structure and lowered the viscosity. Liao et al. [7], Zhang et al. [8], and Yan et al. [9] also investigated the viscosities of BF slag with varying TiO2 content, and the results showed TiO2 reduced the apparent viscosity of molten slags. Zhang et al. [10] revealed the influence of the MgO/Al2O3 ratio on the viscosity of high-Al BF slag and recommended the proper ratio in a range of 0.6–0.7. Although the effect of TiO2 and the MgO/Al2O3 ratio on the viscosity of BF slags at a constant temperature has been investigated, limited research has been conducted at a fixed heat quantity supply. During the BF operation, the fuel ratio and furnace conditions are basically steady; that is, the heat quantity supply in a blast furnace is steady. However, the composition of slag affects its heat capacity, which leads to the different temperature and viscous behavior of slag. Hence, it is of great importance to investigate the relation between the thermodynamic properties of slag with different compositions and its viscosity.
In this study, the viscosity of BF slag with varying TiO2 content and MgO/Al2O3 ratios was measured. According to the thermodynamic data, the heat capacity and enthalpy change of slag were calculated. In addition, the variation of the slag structure with different TiO2 contents was analyzed using Fourier Transform Infrared (FTIR) spectroscopy.

2. Materials and Methods

2.1. Thermodynamic Calculations

The compositions of experimental slags at fixed basicity (C/S = 1.10) are listed in Table 1, and the mass of each sample in the process of calculations was set as 100 g. Factsage thermodynamic software (version 7.1, CRCT & GTT, Montreal, QC, Canada) was used to calculate the heat capacity and enthalpy change of slag, and then the results were corrected according to Equations (1)–(9) to eliminate the unavoidable influence of chemical reaction, crystal transition, and phase change heat. In addition, the liquid temperature of each group of the slags was also calculated using Factsage to ensure that the heat capacity and enthalpy change were calculated in the complete liquid phase region. Factsage is one of the largest computing systems in the chemical thermodynamics field and has been widely used to calculate the thermodynamic data of molten slag [11,12,13].
In the process of calculations, the initial conditions were set as 298 K and 101,325 Pa. The heat capacity and enthalpy change of slag at 1723, 1773, 1823, and 1873 K were calculated. Theoretically, when the pressure of a system is constant and the non-volume work is not conducted, the enthalpy change of slag is approximately equal to the heat absorbed by slag, which is known from the first law of thermodynamics. The average enthalpy changes of slag with varying TiO2 content and MgO/Al2O3 ratios at 1773 K and 1823 K were calculated to use as the given heat quantity supply, and then the temperature of slag at the corresponding heat input was calculated.
CpCaO = 1.048 − 2.046 × 104 T−2 − 2.388 T−1/2 + 1.836 × 106 T−3 (298 K to 2845 K)
CpSiO2 = 1.332 − 5.903 × 104 T−2 − 3.999 T−1/2 + 8.181 × 106 T−3 (298 K to 1996 K)
CpMgO = 1.516 − 1.541 × 104 T−2 − 7.349 T−1/2 + 1.45 × 104 T−3 (298 K to 3098 K)
CpAl2O3 = −0.1541 − 2.973 × 10−4T − 4.899 × 104 T−2 − 1.099 × 103 T−1 + 69.33T−1/2 (298 K to 1200 K)
CpAl2O3 = −7.724 + 6.461 × 10−4 T + 2.587 × 106 T−2 − 1.496 × 104 T−1 + 6.56 × 102 T−1/2 (1200 K to 2327 K)
CpTiO2 = 0.975 − 4.217 × 104 T−2 + 5.045 × 106 T−3 (298 K to 2130 K)
C p   =   m i C pi
Δ H i   =   298 T tr C pi dT   +   Δ tr Hi   +   T tr T M C Pi s dT   +   Δ s l Hi   +   T M T C Pi l dT
Δ H T   =   m i Δ H i
where Cpi is the specific heat capacities of component i, J/g·K; i is the slag components of CaO, SiO2, MgO, Al2O3, and TiO2, J/g·K; Cp is the thermal capacity of the slag at a certain temperature J/K; mi is the mass of component i, g; ΔHi is the enthalpy change of component i at the target temperature; Δ tr Hi is the crystal transition enthalpy of component i at the target temperature, J/g; Δ s l Hi is the melting enthalpy of component i at the given temperature, J/g; Δ H T is the enthalpy change of slag at a given temperature, kJ; Ttr and TM are the crystal transition temperature and melting temperature, respectively, K.

2.2. Experimental Materials

The slags were prepared with the analytical reagents CaO, SiO2, MgO, Al2O3, and TiO2 based on the design of slag. All the reagents were heated in an electric resistance furnace at 1273 K for 150 min to remove impurities. The mixture was precisely weighed at 180 g and packed into a Mo crucible. The samples were heated to 1873 K in a muffle furnace with high-purity argon gas (99.99%, 0.5 L/min). After 120 min heat preservation, the slag was rapidly quenched in distilled water, and the cooled slag sample was taken out and dried. All quenched slags were colorless and transparent, which suggested the main valence state of Ti ions is Ti4+. The quenched slag was ground and then analyzed by X-ray diffraction (XRD, MAXima-7000, Shimadzu, Kyoto, Japan) that used Cu Kα radiation (λ = 1.5406 Å) with the scan rate of 6°/min in the 2 degree range of 10° to 90°. Figure 1 shows the XRD pattern of the representative slag samples. It could be seen that there were no characteristic peaks, which proved that the quenched slag was amorphous. FTIR spectroscopy (Nicolet 6700; Thermo Fisher Scientific, Waltham, MA, USA) was used to analyze the slag structure. During the measurement, 3.0 mg quenched slag powders and 350 mg KBr were mixed and pressed into a film. The scanning time was set as 32 s. The wavenumber range of 400–4000 cm−1 was recorded with a resolution of 4 cm−1.

2.3. Viscosity Measurement

The viscosity measurements were conducted with the rotating spindle method. The experimental apparatus used to measure the viscosity of slags was shown in Figure 2. Using seven MoSi2 heat elements to heat the furnace, a Pt–6%Rh and Pt–30%Rh thermocouple was used to monitor and control the furnace temperature (error ± 2 K). Then, the computer calculated and recorded the viscosity values when the spindle was rotating.
The Mo crucible filled with 110 g of pre-melted slag was put at the constant temperature zone of the furnace. Then, the temperature of the furnace was heated to 1873 K under the protection of argon gas and held for 90 min to ensure the slag was in the fully liquid phase. The spindle was carefully adjusted to the central axis line of the crucible and immersed into the fluid slag. Viscosity measurements were carried out at every 10 K from 1873 K. In order to ensure sufficient thermal equilibration, the furnace was held for 10 min before every measurement. The measurements were ended when the viscosity value of the melt was > 5 Pa·s. The average of three test results was used as the final viscosity value.

3. Results and Discussion

3.1. Effect of TiO2 on Viscosity, Heat Capacity and Enthalpy Change of Slag

Figure 3 graphically represents the viscosity of slag with different TiO2 contents at constant basicity of 1.10. It could be seen from Figure 3 that the slag viscosity decreased with the increment in temperature and TiO2 content. TiO2 had little effect on the decrease in viscosity at experimental temperatures, which is supposed to be due to the intricate network structure that has been depolymerized into simple units under high temperature [14,15]. The variation trends of viscosity are consistent with many previous works on different slag systems. Zheng et al. [16] investigated the effect of TiO2 content (0–30 wt.%) on the viscosity of CaO–SiO2–TiO2 slag at fixed basicity (C/S = 1.2) and concluded that the TiO2 additions weakened the strength of the silicate network and then reduced the viscosity. Jiao et al. [17] showed that the viscosity of CaO–MgO–Al2O3–SiO2–TiO2–FeO BF primary slags decreased with increasing TiO2. It is generally considered that the variation in slag viscosity is closely related to the slag structure. Thus, the decrement in viscosity might be attributed to the existence of numerous silicates or silicate-aluminate structural units in the present melts, and TiO2 may break up the 3D networks consisting of Si–O or Al–O, which results in decreasing the degree of polymerization (DOP) of slag.
The heat capacity and enthalpy change of slag with varying TiO2 contents are shown in Figure 4. It could be observed from Figure 4a that the heat capacity of slag presented a linear decreasing trend as the TiO2 content increased at the target temperature, and it also decreased as the temperature increased at a constant TiO2 content. Meanwhile, when the temperature rose from 1723 K to 1873 K, the decrease rate of heat capacity augmented with increase in TiO2 content. Figure 4b shows the enthalpy change of slag as a function of TiO2 content at different temperatures. It could be seen that the enthalpy change decreased slightly with the increase in TiO2 content and increased dramatically with increasing temperature. The heat capacity and enthalpy change of slag are the important physicochemical properties of slag. The higher the heat capacity, the more heat energy slag needs to absorb when its temperature rises by one degree Celsius, which means that the temperature of slag with a larger heat capacity is difficult to change as the heat quantity supply varies. Therefore, the low TiO2 content in slag is beneficial to keep the temperature of slag constant. As mentioned earlier, the heat absorbed by the slag is approximately equal to the enthalpy change value of the slag, so the increase in TiO2 content and the decrease in temperature could decrease the heat storage capacity of slag. It was suggested that when the TiO2 content increased in the range of 12 to 20 mass%, the heat supply of the BF should maintain stability due to the large fluctuation of slag temperature.
The effects of TiO2 on the viscosity and temperature of slag at the fixed heat quantity supply are depicted in Figure 5. The value of constant heat input was determined by the average value of the enthalpy change of slag with varying TiO2 content at 1773 K and 1823 K. The calculated heat quantity supplies were 155,601 J and 161,289 J, respectively. An increment in the TiO2 content led to higher slag temperature and lower slag viscosity. Due to the heat storage capacity of slag decreasing as TiO2 content increased, the increase in residual heat raised the slag temperature, which had a significant influence on the network structure of slag [14]. By comparing Figure 3 and Figure 5, it also could be observed that the viscosity under the constant heat quantity supply changed relatively greater than the one at the fixed temperature. Therefore, when the TiO2 content in BF slag increases, it is essential to maintain a steady heat quantity input to avoid greater fluctuations in slag temperature and fluidity.

3.2. Effect of the MgO/Al2O3 Ratio on Viscosity, Heat Capacity, and Enthalpy Change of Slag

Figure 6 shows the effect of MgO/Al2O3 ratios on the viscosity of slag at different temperatures. The increase in MgO/Al2O3 ratios lowered the viscosity of slag at the experimental temperature, and the decrement in viscosity at high temperatures was not obvious. This is consistent with the previous studies [10,18,19,20], where MgO is classified as a basic oxide similar to CaO in the silicate melts, and it acts as a network modifier. Furthermore, MgO could provide free oxygen ions to cut down the Si–O bonds in the network structure, resulting in the decrease in the DOP of the slag structure, expressed as Equations (10) and (11) [4,21]. These transition processes could improve the fluidity of molten slag.
[Si3O9]6− (ring) + O2− = [Si3O10]8− (chain)
[Si3O10]8− + O2− = [Si2O7]6− + [SiO4]4− (monomer)
The effects of the MgO/Al2O3 ratio on the heat capacity and enthalpy change of slag at various temperatures are depicted in Figure 7. It shows that the increment in MgO/Al2O3 led to a larger heat capacity of slag at the constant temperature, and the increment in heat capacity was much larger than that caused by the addition of TiO2. The enthalpy changes of slag also increased with increasing MgO/Al2O3 ratios and temperature. However, the absolute values of the enthalpy change at the experimental temperature only increased slightly with increasing MgO/Al2O3 ratios, which was basically consistent with the previous study. It was reported in the work [19] that a proper increase in the MgO/Al2O3 ratio was beneficial to reduce the fuel ratio or coke ratio of the blast furnace to some extent.
Figure 8 shows the viscosity and temperature of slag as a function of MgO/Al2O3 ratios under fixed heat input conditions. The calculated heat input quantities were 155,042 J (1773 K) and 160,593 J (1823 K), respectively. It could be seen that the temperature in slag decreased with increasing MgO/Al2O3 ratios, and the average decrement of slag temperature was 5 K. However, it is interesting to note that although the temperature of slag decreased, the viscosity had not increased but decreased. This showed that the influence of the slag composition on the viscosity was more pronounced than the slight decrease in temperature under a fixed heat supply. Therefore, the energy consumption of the BF could be reduced by appropriately increasing the MgO/Al2O3 ratio of slag.

3.3. Effect of TiO2 and MgO/Al2O3 Ratio on the Activation Energy for Viscous Flow

Generally, the activation energy represents the energy barrier to overcome when the liquid melt flows, and the variation in activation energy reflects the changes to slag structures. This can be calculated from the Arrhenius-type equation [22,23], as expressed by Equation (12).
η   =   AT   exp   ( E a   RT   )
Equation (12) can be written as Equation (13) after taking the logarithm.
ln ( η T )   =   lnA   +   E a RT
where η is the viscosity of the melt, (Pa·s); A is the pre-exponent factor, and Ea is the apparent activation energy of the slags, (J/mol); R is the gas constant, (8.314 J/mol·K); T is the absolute temperature, (K).
From Equation (4), ln( η T ) has a linear relationship with 1 RT , which is shown in Figure 9. The slope of the fitting line (R2 > 0.95) is the apparent activation energy Ea. The calculated results of apparent activation energy are listed in Table 2. As expected, the apparent activation energy decreased from 123.89 kJ/mol to 117.67 kJ/mol with the TiO2 content from 12 to 20 mass%, and it also decreased from 121.61 kJ/mol to 115.96 kJ/mol with the increment of the MgO/Al2O3 ratio from 0.4 to 0.8. This indicated that the frictional resistance of melts and the fraction of complex structural units decreased. The results were consistent with the variation of viscosity.

3.4. Effect of TiO2 on the Structure Using FTIR Spectroscopy

The FTIR spectra of the quenched slag were measured to obtain information about the slag structure. The original FTIR spectra for the amorphous samples with different TiO2 content at fixed basicity of 1.10 are presented in Figure 10. According to the previous studies [24,25,26], the vibration bands of silicate are assigned to the range of 1200 to 400 cm−1. The spectra were divided into three wavenumber bands: 1200–760 cm−1, 760–600 cm−1, and 600–400 cm−1. For [SiO4]-tetrahedral bands, it was divided into four typical silicate structural units Qn (n = 0, 1, 2, 3, which represents the number of bridge oxygen atoms per silicate tetrahedral structure) and represented the DOP of the melted structure. The assignment of each band is shown in Table 3.
From Figure 10 and Table 3, as the TiO2 content increased, the trough of [SiO4]-tetrahedral symmetric stretching bands located at 800–1200 cm−1 became less pronounced, and the depth of the transmission trough got shallower. Furthermore, the band center of the [SiO4]-tetrahedral shifted from 996.37 cm−1 to 965.32 cm−1 as the TiO2 content increased. This indicated that the intricate three-dimensional silicate network structures were depolymerized and formed simpler structural units. In addition, the asymmetric stretching vibration bands for the [AlO4]-tetrahedron dampened with increasing TiO2 as well, and the wavenumber of [AlO4]-tetrahedron bands gradually decreased from 764.62 cm−1 to 760.50 cm−1. It suggested that TiO2 also played a role in depolymerizing the aluminate structure. According to Xu et al. [31], Ti4+ might incorporate into the slag structure and substitute some Al3+ in [AlO4]5− to form [TiO4]4−, but cations (e.g., Ca2+) are needed to keep the equilibrium of electric charge. When Ti4+ substituted Al3+, [TiO4]4− generated and [AlO4]5− disappeared, which meant less cations consumption. Thus, the oxide (e.g., CaO) acted as network modifier, accordingly increased, and viscosity decreased. The depth of the T–O–T (T donates Si, Al, or Ti) bending band significantly decreased with increasing TiO2 content. This further illustrated a decrease in the linkage between [SiO4]-tetrahedron and [SiO4]-tetrahedron or [SiO4]-tetrahedron and [AlO4]-tetrahedron, and the complex structural units were depolymerized. This result is because [TiO4]-tetrahedral units entered melts and incorporated silicate structures, and Si–O–Ti units were formed according to Equation (14). Since the electrostatic potential of Ti4+ (1.85 I) is lower than that of Si4+ (2.51 I), the bond strength of Ti–O is weaker than Si–O, and the stability of slag weakens [32,33]. In addition, according to the stable energy concept proposed by Duan et al. [34], the stable energies of [TiO4] and [SiO4] were calculated to be approximately 7.282 × 10−5 and 8.426 × 10−5 kJ·mol−1, respectively. It suggested that [TiO4] was more easily destroyed than [SiO4], which meant that the slag needed to absorb less heat energy, that is, the heat capacity of slag decreased. Therefore, the heat capacity and viscosity of slag decreased. The reduction in viscosity indicated that TiO2 played the role of a network modifier to weaken the complexity of the network structure. The comprehensive effects of TiO2 may eventually cause the decrease in slag viscosity.
Si−O−Si + 2[TiO4] → 2Si−O−Ti

4. Conclusions

The effects of TiO2 content (12 mass% to 20 mass%) and MgO/Al2O3 (0.4 to 0.8) ratios on the viscosity, heat capacity, and enthalpy change of Ti-bearing BF slag at a constant heat supply were investigated, and the structure of slag with varying TiO2 contents was studied by using FTIR spectroscopy. The following conclusions could be drawn.
The viscosity of slag decreased as the TiO2 content and MgO/Al2O3 ratio increased at a constant temperature, and the heat capacity and enthalpy change both decreased with increasing TiO2 content. With the increase in the MgO/Al2O3 ratio, the heat capacity of slag increased, and the absolute values of the enthalpy change only increased slightly at the experimental temperature.
  • At the fixed heat quantity supply, the increase in TiO2 content led to an increase in slag temperature but a decrease in the viscosity. As the MgO/Al2O3 ratio increased, the temperature of slag had the opposite variation trend, but the viscosity of slag still decreased.
  • The apparent activation energy decreased from 123.89 kJ/mol to 117.67 kJ/mol with the increment of TiO2 content, and it also decreased from 121.61 kJ/mol to 115.96 kJ/mol with higher MgO/Al2O3 ratios.
  • The vibration bands of [SiO4]-tetrahedron, [AlO4]-tetrahedron, and T–O–T became less pronounced. TiO2 played a role in weakening the stability of the molten slag structure and depolymerizing the complex network structural units into simpler structures, which eventually decreased the viscosity.

Author Contributions

Conceptualization, X.X. and Z.P.; methodology, X.X. and Z.P.; software, J.F., and R.X.; validation, X.X. and Z.P.; formal analysis, J.F. and X.X.; investigation, X.X.; resources, J.F., X.X. and R.X.; data curation, X.X., J.F. and Z.P.; writing—original draft preparation, X.X. and Z.P.; writing—review and editing, X.X.; supervision, X.X. and Z.P.; project administration, X.X.; funding acquisition, X.X. All authors have read and agreed to the published version of the manuscript.

Funding

The present work was financially supported by the China Postdoctoral Science Found (Grant No. 2019M663932XB) and Natural Science Basic Research Program of Shaanxi (Program No. 2019JLP-05). The authors gratefully acknowledge their support.

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. Jiao, K.; Zhang, J.; Chen, C.; Liu, Z.; Jiang, X. Operation characteristic of super-large blast furnace slag in China. ISIJ Int. 2017, 57, 983–998. [Google Scholar]
  2. Husslage, W.M.; Bakker, T.; Steeghs, A.G.S.; Reuter, M.A.; Heerema, R.H. Flow of molten slag and iron at 1500 °C to 1600 °C through packed coke beds. Metall. Mater. Trans. B 2005, 36, 765–776. [Google Scholar] [CrossRef]
  3. Gao, K.; Jiao, K.; Zhang, J.; Dianyu, E. Thermodynamic properties and viscosities of high-titanium slags. ISIJ Int. 2020, 60, 1902–1908. [Google Scholar] [CrossRef]
  4. Xing, X.; Pang, Z.; Zheng, J.; Du, Y.; Ren, S.; Ju, J. Effect of MgO and K2O on high-Al silicon–manganese alloy slag viscosity and structure. Minerals 2020, 10, 810. [Google Scholar] [CrossRef]
  5. Hyunsik, P.; Park, J.Y.; Hyun, K.; 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]
  6. Sohn, I.; Wang, W.; Hiroyuki, M.; Fumitaka, T.; Dong, 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]
  7. 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]
  8. Zhang, G.; Chou, K.; Zhang, J. Influence of TiO2 on viscosity of aluminosilicate melts. Ironmak. Steelmak. 2014, 41, 47–50. [Google Scholar] [CrossRef]
  9. Yan, Z.; Lv, X.; He, W.; Xu, J. Effect of TiO2 on the liquid zone and apparent viscosity of SiO2–CaO–8 wt.%MgO–14 wt.%Al2O3 system. ISIJ Int. 2017, 57, 31–36. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, X.; Jiang, T.; Xue, X.; Hu, B. Influence of MgO/Al2O3 ratio on the viscosity of blast furnace slag with high Al2O3 content. Steel Res. Int. 2016, 87, 87–93. [Google Scholar] [CrossRef]
  11. Kou, M.; Wu, S.; Ma, X.; Wang, L.; Chen, M.; Cai, Q.; Zhao, B. Phase equilibrium studies of CaO-SiO2-MgO-Al2O3 system with binary basicity of 1.5 related to blast furnace slag. Metall. Mater. Trans. B 2016, 47, 1093–1102. [Google Scholar] [CrossRef]
  12. Suzuki, M.; Jak, E. Quasi-chemical viscosity model for fully liquid slag in the Al2O3–CaO–MgO–SiO2 system–Part I: Revision of the model. Metall. Mater. Trans. B 2013, 44, 1435–1450. [Google Scholar] [CrossRef] [Green Version]
  13. Suzuki, M.; Jak, E. Quasi-chemical viscosity model for fully liquid slag in the Al2O3–CaO–MgO–SiO2 system–Part II: Evaluation of slag viscosities. Metall. Mater. Trans. B 2013, 44, 1451–1465. [Google Scholar] [CrossRef] [Green Version]
  14. Chang, Z.Y.; Jiao, K.X.; Zhang, J.-L.; Ning, X.-J.; Liu, Z.Q. Effect of TiO2 and MnO on viscosity of blast furnace slag and thermodynamic analysis. ISIJ Int. 2018, 58, 2173–2179. [Google Scholar] [CrossRef] [Green Version]
  15. Kim, J.; Sohn, I. Influence of TiO2/SiO2 and MnO on the viscosity and structure in the TiO2–MnO–SiO2 welding flux system. J. Non-Cryst. Solids 2013, 359, 235–243. [Google Scholar] [CrossRef]
  16. 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]
  17. Jiao, K.X.; Zhang, J.-L.; Wang, Z.Y.; Chen, C.-H.; Liu, Y.X. Effect of TiO2 and FeO on the viscosity and structure of blast furnace primary slags. Steel Res. Int. 2017, 88, 1600296. [Google Scholar] [CrossRef]
  18. Li, P.; Ning, X. Effects of MgO/Al2O3 ratio and basicity on the viscosities of CaO–MgO–SiO2–Al2O3 slags: Experiments and modeling. Metall. Mater. Trans. B 2016, 47, 446–457. [Google Scholar]
  19. Jiao, K.; Chang, Z.-Y.; Chen, C.; Zhang, J.-L. Thermodynamic properties and viscosities of CaO–SiO2–MgO–Al2O3 slags. Metall. Mater. Trans. B 2019, 50, 1012–1022. [Google Scholar] [CrossRef]
  20. Zhang, S.; Zhang, X.; Liu, W.; Lv, X.; Bai, C.; 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]
  21. Park, J.H.; Min, D.J.; Song, H.S. FT-IR spectroscopic study on structure of CaO–SiO2 and CaO–SiO2–CaF2 slags. ISIJ Int. 2002, 42, 344–351. [Google Scholar] [CrossRef]
  22. Hu, K.; Lv, X.; Li, S.; Lv, W.; Song, B.; Han, K. Viscosity of TiO2–FeO–Ti2O3–SiO2–MgO–CaO–Al2O3 for high-titania slag smelting process. Metall. Mater. Trans. B 2018, 49, 1963–1973. [Google Scholar] [CrossRef]
  23. Qi, J.; Liu, C.; Jiang, M. Role of Li2O on the structure and viscosity in CaO–Al2O3–Li2O–Ce2O3 melts. J. Non-Cryst. Solids 2017, 475, 101–107. [Google Scholar] [CrossRef]
  24. Mysen, B.; Ryerson, F.; Virgo, D. The Influence of TiO2 on the structure and derivative properties of silicate melts. Am. Miner. 1980, 68, 1150–1165. [Google Scholar]
  25. Seung, M.; Park, J.; Sohn, I. Surface kinetics of nitrogen dissolution and its correlation to the slag structure in the CaO-SiO2, CaO-Al2O3, and CaO-SiO2-Al2O3 slag system. J. Non-Cryst. Solids 2011, 357, 2868–2875. [Google Scholar]
  26. Jiang, Y.; Lin, X.; Ideta, K.; Takebe, H.; Miyawaki, J.; Yoon, S.-H.; Mochida, I. Microstructural transformations of two representative slags at high temperatures and effects on the viscosity. J. Ind. Eng. Chem. 2014, 20, 1338–1345. [Google Scholar] [CrossRef]
  27. Wang, W.; Dai, S.; Zhou, L.-J.; Zhang, J.; Tian, W.; Xu, J. Viscosity and Structure of MgO–SiO2-based Slag Melt with varying B2O3 Content. Ceram. Int. 2020, 46, 3631–3636. [Google Scholar] [CrossRef]
  28. Sun, Y.; Liao, J.; Zheng, K.; Wang, X.; Zhang, Z. Effect of B2O3 on the structure and viscous behavior of Ti-bearing blast furnace slags. JOM 2014, 66, 2168–2175. [Google Scholar] [CrossRef]
  29. Wang, H.; Chen, Z.; Liu, L.; Sun, Y.; Wang, X. Roles of P2O5 Addition on the viscosity and structure of CaO–SiO2–Al2O3–Na2O–P2O5 melts. ISIJ Int. 2018, 58, 1644–1649. [Google Scholar] [CrossRef]
  30. Kim, H.; Matsuura, H.; Tsukihashi, F.; Wang, W.; Min, D.; 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 2013, 44, 5–12. [Google Scholar] [CrossRef]
  31. Xu, R.; Zhang, J.; Ma, R.; Jiao, K.; Zhao, Y. Influence of TiO2 on the viscosity of a high alumina slag and on carbon brick corrosion. Steel Res. Int. 2017, 1700353. [Google Scholar] [CrossRef]
  32. Huang, W.; Zhao, Y.; Yu, S.; Zhang, L.; Ye, Z.; Wang, N.; Chen, M. Viscosity property and structure analysis of FeO–SiO2–V2O3–TiO2–Cr2O3 slags. ISIJ Int. 2016, 56, 594–601. [Google Scholar] [CrossRef] [Green Version]
  33. Wang, Z.; Shu, Q.; Chou, K. Viscosity of fluoride-free mold fluxes containing B2O3 and TiO2. Steel Res. Int. 2013, 84, 766–776. [Google Scholar] [CrossRef]
  34. Duan, R.-G.; Liang, K.-M.; Gu, S.-R. The stable energy of glass structure unit. Mater. Trans. JIM. 1998, 39, 1162–1163. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The XRD pattern of CaO–SiO2–Al2O3–CaO–MgO–TiO2 slag.
Figure 1. The XRD pattern of CaO–SiO2–Al2O3–CaO–MgO–TiO2 slag.
Materials 14 00124 g001
Figure 2. Experimental apparatus for the measurement of slag viscosity.
Figure 2. Experimental apparatus for the measurement of slag viscosity.
Materials 14 00124 g002
Figure 3. The viscosity of blast furnace (BF) slag with different TiO2 contents.
Figure 3. The viscosity of blast furnace (BF) slag with different TiO2 contents.
Materials 14 00124 g003
Figure 4. The heat capacity and enthalpy change of slag with varying TiO2 contents; (a) heat capacity; (b) enthalpy change.
Figure 4. The heat capacity and enthalpy change of slag with varying TiO2 contents; (a) heat capacity; (b) enthalpy change.
Materials 14 00124 g004
Figure 5. The viscosity of slag with varying TiO2 content under constant heat quantity: (a) 155,601 J; (b) 161,289 J.
Figure 5. The viscosity of slag with varying TiO2 content under constant heat quantity: (a) 155,601 J; (b) 161,289 J.
Materials 14 00124 g005
Figure 6. The viscosity of BF slag with different MgO/Al2O3 ratios.
Figure 6. The viscosity of BF slag with different MgO/Al2O3 ratios.
Materials 14 00124 g006
Figure 7. The heat capacity and enthalpy change of slag with different MgO/Al2O3 ratios; (a) heat capacity; (b) enthalpy change.
Figure 7. The heat capacity and enthalpy change of slag with different MgO/Al2O3 ratios; (a) heat capacity; (b) enthalpy change.
Materials 14 00124 g007
Figure 8. The viscosity of slag with different MgO/Al2O3 ratios under constant heat quantity: (a) 155,042 J; (b) 160,593 J.
Figure 8. The viscosity of slag with different MgO/Al2O3 ratios under constant heat quantity: (a) 155,042 J; (b) 160,593 J.
Materials 14 00124 g008
Figure 9. Temperature dependence of the viscosity of the slags (a) with varying TiO2 content; (b) with varying MgO/Al2O3 ratio.
Figure 9. Temperature dependence of the viscosity of the slags (a) with varying TiO2 content; (b) with varying MgO/Al2O3 ratio.
Materials 14 00124 g009
Figure 10. Effect of TiO2 on the results of FTIR spectra of as-quenched samples.
Figure 10. Effect of TiO2 on the results of FTIR spectra of as-quenched samples.
Materials 14 00124 g010
Table 1. Experimental compositions of slags (mass%).
Table 1. Experimental compositions of slags (mass%).
No.CaOSiO2Al2O3MgOTiO2MgO/Al2O3Liquid Temperature (K)
130.0327.3014.178.50200.61432
231.0828.2514.178.50181430
332.1329.2014.178.50161427
433.1730.1614.178.50141421
534.2231.1114.178.50121413
630.0327.3016.196.48200.41440
730.0327.3015.117.56200.51436
830.0327.3013.349.33200.71428
930.0327.3012.5910.08200.81424
Table 2. The calculated results of apparent activation energy.
Table 2. The calculated results of apparent activation energy.
TiO2 (mass%)Ea (kJ/mol)MgO/Al2O3
(mass%/mass%)
Ea (kJ/mol)
20117.670.4121.61
18119.390.5118.18
16120.320.6117.67
14122.420.7116.27
12123.890.8115.96
Table 3. Assignments of FTIR spectra bands associated with structural units.
Table 3. Assignments of FTIR spectra bands associated with structural units.
Wavenumber (cm−1)AssignmentsReferences
400–600T-O-T bond bending vibrations (T denotes Si, Al, or Ti elements, etc.)[17,18,27,28]
600–760The symmetric stretching vibration of [AlO4]-tetrahedron[14,18,29,30]
~850Q0 ([SiO4]4 monomers)[14,16,29,30]
~900Q1 ([Si2O7]6 dimers)[14,16,29,30]
~950–1000Q2 ([Si3O10]8 chains)[14,16,29,30]
~1050Q3 ([Si3O9]6 rings)[14,16,29,30]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fang, J.; Pang, Z.; Xing, X.; Xu, R. Thermodynamic Properties, Viscosity, and Structure of CaO–SiO2–MgO–Al2O3–TiO2–Based Slag. Materials 2021, 14, 124. https://doi.org/10.3390/ma14010124

AMA Style

Fang J, Pang Z, Xing X, Xu R. Thermodynamic Properties, Viscosity, and Structure of CaO–SiO2–MgO–Al2O3–TiO2–Based Slag. Materials. 2021; 14(1):124. https://doi.org/10.3390/ma14010124

Chicago/Turabian Style

Fang, Jinle, Zhuogang Pang, Xiangdong Xing, and Runsheng Xu. 2021. "Thermodynamic Properties, Viscosity, and Structure of CaO–SiO2–MgO–Al2O3–TiO2–Based Slag" Materials 14, no. 1: 124. https://doi.org/10.3390/ma14010124

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