ISIJ International
Online ISSN : 1347-5460
Print ISSN : 0915-1559
ISSN-L : 0915-1559
Regular Article
Selective Phase Transformation Behavior of Titanium-bearing Electric Furnace Molten Slag during the Molten NaOH Treatment Process
Yang LiHai-yang YuZuo-tai ZhangMei ZhangMin Guo
Author information
JOURNAL OPEN ACCESS FULL-TEXT HTML

2015 Volume 55 Issue 1 Pages 134-141

Details
Abstract

In this paper, selective phase transformation for titanium-bearing phase (Ti-bearing phase) and impurity phase in titanium-bearing electric arc furnace molten slag (Ti-bearing EAF slag) was effectively realized during the molten NaOH treatment process. The phase transformation mechanism based on thermodynamic calculation was investigated in detail by X-ray diffraction, X-ray photoelectron spectroscopy and Raman spectrum. It is indicated that the Ti-bearing phases in Ti-bearing EAF slag such as anosovite solid solution and Mg2TiO4 can be easily converted to Na2TiO3 with NaCl-type crystal structure, whilst the main impurity phase such as MgAl2O4 was apt to be decomposed by molten NaOH to form NaAlO2 and MgO. In addition, with increasing of roasting temperature and time, as well as decreasing of Ti-bearing EAF slag/NaOH mass ratio (Rslag/NaOH), the formed Na2TiO3 could be partly changed to NaMO2 (M=Mg, Ti, Fe) with α-NaFeO2-type crystal structure, due to the coexisting metal ions like Mg2+ and Fe3+ in the slag be doped into or even substituted the Ti atoms of Na2TiO3 to form NaMO2 (M=Ti, Fe, Mg).

1. Introduction

Ti-bearing slags such as Ti-bearing blast furnace slag (Ti-bearing BF slag) and Ti-bearing electric arc furnace molten slag (Ti-bearing EAF slag) etc., are considered as valuable secondary resources in China due to their relatively higher TiO2 contents. Since the Ti-bearing slags are produced from various raw ores by using different treating processes, the main chemical compositions and the phase structure of these slags are basically different. Therefore, how to extract titanium efficiently from the Ti-bearing slags has been attracted much attention, and various approaches have been utilized including acid process,1,2,3) alkali fusion process,4,5) carbonization and chlorination process,6) selective crystallization process.7,8,9,10,11) Comparing these methods, the alkali fusion process shows obvious advantage such as unrestricted raw slag materials, and easy recycling of alkali, etc. More importantly, after alkali fusion process, the Ti-bearing phases can be selectively separated from other impurity phases by only water leaching process based on the different water solubility of formed sodium salts instead of complicate dressing method.

However, no matter which Ti-bearing slags were used during the alkali fusion process, the attention of researchers was mainly focused on the extraction efficiency of titanium and the decomposition kinetics of Ti-bearing slags in different alkali systems. None of the efforts have been taken on phase transformation mechanism for titanium and other elements in slags. Zhang et al.12) studied the decomposition behavior of Ti-bearing slag in NaOH solution at 493 K for 4 h, and indicated that the hydrothermal product obtained under optimal conditions was Na4Ti3O8. Scott Middlemas et al.13) believed that the Ti-bearing phases in Ti-bearing slag could react with molten NaOH at 773 K for 4 h to form Na2TiO3. Han et al.5) also considered that Ti-bearing slag reacted with molten NaOH yielded only one kind of sodium titanate (Na2TiO3). In fact, it should be noted that the titania could react with molten NaOH to form various sodium salts.14,15,16,17) However, the phase transformation mechanism of the Ti-bearing phases was not clear, especially the influence of metal ions such as Fe3+, Al3+, Mg2+, etc. coexisting in the slags on the formation of Ti-bearing phases were seldom been investigated.

In this paper, Ti-bearing EAF slag was used as raw material to investigate the transformation mechanism of the Ti-bearing phases and impurity phases during the molten NaOH treatment process. The phase transformation mechanism was studied in detail based on thermodynamic calculation and crystal structure analysis. This paper may provide some essential thermodynamic information for Ti-bearing slags or other Ti-bearing materials in the molten NaOH system, and would also give a reference to comprehensive utilization of titanium resource treated by molten NaOH treatment process.

2. Experimental

2.1. Materials and Characterization

All the chemical reagents employed were analytical grade (Sinopharm Chemical Reagent Co. Ltd) and the distilled water was used throughout the experiment. The Ti-bearing EAF slag samples were provided by Panzhihua Steel Company (Sichuan Province, China), which were obtained from vanadium-titanium magnetite concentrate through direct reduction reaction by rotary hearth furnace and smelting separation by electric furnace. The main chemical compositions of the slag were analyzed by inductively coupled plasma optical emission spectroscopy (ICP-OES, TELEDYNE Leeman Labs) and listed in Table 1. The phase structure of the slag was investigated by X-ray diffraction using Cu Kα radiation (λ = 0.154056 nm) with 40 kV, 200 mA and a speed of 10°/min (XRD, M21X, MAC SCIENCE Co. Ltd, Japan). The results shown in Fig. 1 indicated that the Ti-bearing phases of the slag were MgTi2O5, Ti3O5 and Mg2TiO4, while the main impurity phases were MgAl2O4 and amorphous CaSiO3. The valences of transition elements were analyzed by X-ray photoelectron spectroscopy (XPS, Al Kα, AXIS ULTRA), in which C1s binding energy value (284.8 eV) was chosen as standard for baseline correction and baseline subtraction method was Shirley-type. The XPS spectrum of Ti in Ti-bearing EAF slag was shown in Fig. 2. It is shown that about 15% Ti3+ and 85% Ti4+ coexisted in the slag, suggesting that the existence form of titanium element might be Ti3O5 or Ti2O3 anosovite solid solutions in the slag.

Table 1. Main chemical compositions of the Ti-bearing EAF slag (mass%).
CompositionTiO2Al2O3MgOSiO2CaOFe2O3
Content50.919.412.98.05.42.9
Fig. 1.

XRD pattern of the Ti-bearing EAF slag.

Fig. 2.

XPS spectrum of Ti in the Ti-bearing EAF slag. (Online version in color.)

Measurements of Raman spectra were performed on a laser confocal Raman spectrometer (JY-HR800, Jobin Yvon). A blue line (514 nm) of the laser was taken as the excitation source.

Artificially synthesized MgTi2O5 was prepared by roasting TiO2 (anatase, analytical reagent) with MgO at 1873 K for 3 h with TiO2/MgO molar ratio at 2:1.

2.2. The Concrete Process of Extracting TiO2 from Ti-bearing EAF Slag

Figure 3 illustrated the general flow sheet of whole extraction process of TiO2 from Ti-bearing EAF slag, and the detailed procedure were described as follows.18) It should be noted that this paper was mainly focus on the molten NaOH treatment process to elucidate the Ti-bearing EAF slag phase transformation mechanism clearly.

Fig. 3.

Schematic flow diagram of extracting TiO2 from Ti-bearing EAF slag.

Firstly, 10 g Ti-bearing EAF slag was ground to about 120 mesh and mixed with NaOH homogeneously at different mass ratios in a nickel crucible. Then, the nickel crucible was placed into a muffle furnace when the temperature reached a preset value, holding for a required time with free access to air. After the molten NaOH treatment process, the nickel crucible was taken out rapidly and then cooled at room temperature. The obtained slag was called alkali fusion slag. For phase analysis, the alkali fusion slag was ground to about 200 mesh rapidly.

Secondly, as for hydrolysis reaction process, 3 g ground alkali fusion slag was dissolved in distilled water with electromagnetic stirring for 1 h. The sodium salts such as Na2SiO3 and NaAlO2 formed in the alkali fusion slag were dissolved in water while the insoluble sodium titanate salts also formed in the alkali fusion slag were separated from the solution by filtration.

Finally, the obtained residue was mixed with HCl solution and refluxed at 373 K for 6 h to perform acidolysis reaction to get metatitanic acid. Then, the samples were calcined in the muffle furnace at 773 K for 1 h to obtain TiO2.

3. Results and Discussion

3.1. Thermodynamic Analysis

When the Ti-bearing EAF slag was mixed with NaOH and roasted at different temperatures in muffle furnace under atmospheric condition, the related reactions could be described as follows:   

1 4 MgT i 2 O 5 +NaOH= 1 2 N a 2 Ti O 3 + 1 4 MgO+ 1 2 H 2 O (1)
  
1 6 T i 3 O 5 +NaOH+ 1 12 O 2 = 1 2 N a 2 Ti O 3 + 1 2 H 2 O (2)
  
1 2 M g 2 Ti O 4 +NaOH= 1 2 N a 2 Ti O 3 +MgO+ 1 2 H 2 O (3)
  
1 2 CaSi O 3 +NaOH= 1 2 N a 2 Si O 3 + 1 2 CaO+ H 2 O (4)
  
1 2 MgA l 2 O 4 +NaOH=NaAl O 2 + 1 2 MgO+ 1 2 H 2 O (5)

Therefore, the standard Gibbs free energy change (ΔG°) of the above reactions can be calculated according to Eq. (6).   

Δ G T ° =Δ H T ° -TΔ S T ° =Δ H 298 ° -TΔ S 298 ° + 298 T 1 Δ C p dT+ T 1 T 2 Δ C p dT - T 1 298 T 1 Δ C p T dT- T 2 T 1 T 2 Δ C p T dT (6)

Where ΔG° is the standard Gibbs free energy change, J/mol, ΔH° is the standard enthalpy change, J/mol, ΔS° is the standard entropy change, J/(mol K), Cp is the specific heat capacity at constant pressure, J/(mol K). Thereinto, Δ H 298 ° , S 298 ° and Cp could be obtained from the FactPS database of FactSage 6.3 software, respectively, and the related values were listed in Table A1. Figure 4 gave the variation trend of ΔG° ranging from 573 K to 1173 K. It is shown that the ΔG° of all the reactions were negative and decreased with the roasting temperature increasing, indicating that all the reactions could be occurred above 573 K. Moreover, it should be noted that the ΔG° values of the reactions (1), (2) and (3) were more negative than that of the other reactions, while the ΔG° value of the reaction (5) was most positive among all the reactions, suggesting that the Ti-bearing phases in the slag would react with molten NaOH much earlier than other phases, especially the impurity phase MgAl2O4 during the molten NaOH treatment process based on thermodynamic calculation. Considering that MgAl2O4 was difficult to be decomposed by molten NaOH than other impurity phase, e.g. CaSiO3, so, the investigation on the transformation of the main impurity phases in this paper was only focus on the decomposition of MgAl2O4 during the molten NaOH treatment process.

Fig. 4.

The plot of ΔrG°~T for reactions between Ti-bearing EAF slag and molten NaOH.

Because NaOH can be used as an ionized solvent at higher temperature above 623 K,19) it is reasonable to assume that the reactions between Ti-bearing EAF slag and molten NaOH may be considered as a solid-liquid reaction mechanism. In order to make sure the Ti-bearing phases or other impurity phases in Ti-bearing EAF slag convert to their corresponding sodium salts completely, Ti-bearing EAF slag/ NaOH mass ratio (Rslag/NaOH) should be controlled at least below 1 according to the material balance calculation. Moreover, excess amount of NaOH is necessary to maintain the liquidity of the reactants and ensure their sufficient contact areas during the reactions. Based on the aforementioned analysis, all the experiments should be conducted above 673 K with Rslag/NaOH below 1.

3.2. The Effect of Roasting Temperature on the Phase Transformation

The effect of roasting temperature on the Ti-bearing phases transformation was investigated from 773 K to 973 K for 1 h. Figure 5 gives the XRD patterns of alkaline fusion slag obtained under different temperatures with Rslag/NaOH fixed at 1:1. It can be seen that the diffraction peaks of anosovite solid solution and Mg2TiO4 completely disappeared while strong diffraction peaks of Na2TiO3 (2θ=40.1°) appeared. This case indicated that the anosovite solid solution (Fig. 6(a)) with orthorhombic crystal system was decomposed by molten NaOH and converted to Na2TiO3 (Fig. 6(b)) with cubic crystal structure (until cell parameters of a=b=c=4.49 Å, and α=β=γ=90°). Increasing the temperature to 873 K, a new phase named NaMO2 (M=Ti, Mg, Fe) with hexagonal crystal system (unit cell parameters of a=b=3.04 Å, c=16.26 Å and α=β=90°, γ=120°) appeared, which was derived from α-NaFeO2-type crystal structure (Fig. 6(c)).14) Further increasing the temperature to 973 K, the diffraction peaks of NaMO2 (2θ=16.4°, 41.1°) became stronger, whereas the diffraction peaks of Na2TiO3 (2θ=40.1°) turned to weaker, suggesting that relatively higher temperature (>873 K) may be benefit for Ti-bearing phases (e.g. MgTi2O5 and Mg2TiO4) transformation to NaMO2 instead of Na2TiO3 during the molten NaOH treatment process.

Fig. 5.

XRD patterns of alkaline fusion slags obtained under different temperatures with Rslag/NaOH fixed at 1:1.

Fig. 6.

The phase transformation relationship between Ti-bearing phases (a) MgTi2O5, (b) Na2TiO3, (c) NaMO2. (Online version in color.)

It is known that the compositions of Ti-bearing EAF slag are very complex and various elements including Fe3+ and Mg2+ etc. coexist in the slag system. Therefore, with the roasting temperature increasing, some Fe3+ or Mg2+ ions would be doped into Na2TiO3, or even substituted the Ti atom of Na2TiO3 to form a class of layered ionic structure called α-NaFeO2-type crystal structure.14) In this paper, the M in NaMO2 with α-NaFeO2-type structure phase can be Fe, Mg or Ti elements due to their similar ion radius (Fe3+=0.064 nm, Mg2+=0.066 nm and Ti4+=0.068 nm). With increasing the roasting temperature, more and more Fe3+ or Mg2+ would be doped into Na2TiO3 to make Na2TiO3 convert to NaMO2 (M=Fe3+, Ti4+, Mg2+), so the amount of NaMO2 increased whereas the amount of Na2TiO3 decreased. According to the analysis above, nearly all the Ti-bearing phases in Ti-bearing EAF slag could convert to Na2TiO3 or NaMO2 during the molten NaOH treatment process, implying the Ti-bearing phase transformation can be realized effectively as illustrated in Fig. 6.

Based on the experimental results, one problem need to be elucidated clearly, that is to say, Fe3+ or Mg2+ doping into Na2TiO3 would promote Na2TiO3 to be transformed to NaMO2 (M=Fe3+, Ti4+, Mg2+). In order to confirm the key influence factors on the phase transformation between Na2TiO3 and NaMO2, pure TiO2 (anatase, analytical reagent) was used as raw material to react with NaOH. Meanwhile, other experimental conditions were controlled the same as that during molten NaOH treatment reaction process. Figure 7 shows the XRD patterns of the product obtained at 973 K for 1 h. It can be seen that the main phases of the product were only Na2TiO3 and excess NaOH. None diffraction peaks of NaMO2 appeared, indicating that TiO2 reacted with molten NaOH could not be reduced to form NaTiO2 in air atmosphere.

Fig. 7.

XRD patterns of the product prepared by pure TiO2 with molten NaOH at 973 K for 1 h.

However, as shown in Fig. 5, when Ti-bearing EAF slag was used as raw material to react with molten NaOH, the diffraction peaks of NaMO2 did exist at 873 K and their intensities became stronger with the roasting temperature increasing to 973 K. The results were basically different from each other, suggesting that coexisting elements such as Fe3+ and Mg2+ in the Ti-bearing EAF slag determined the formation of NaMO2 or Na2TiO3. In order to further investigate the transformation mechanism between Na2TiO3 and NaMO2, artificially synthesized MgTi2O5 was used as raw material to react with molten NaOH under the same experimental conditions. Figure 8 gives the XRD patterns of the products obtained under different roasting temperatures. It can be clearly seen that the main phases of the product were Na2TiO3 and MgO when the temperature was fixed at 773 K, while NaMO2 with α-NaFeO2-type crystal structure appeared when the temperature increased to 973 K. So it is reasonable to assume that the synthesized MgTi2O5 could be decomposed by molten NaOH according to the following equations:   

MgT i 2 O 5 +4NaOH2N a 2 Ti O 3 +MgO+2 H 2 O 673 K<T<773 K (7)
  
MgT i 2 O 5 +2NaOH 2Na( M g 1-x , T i x ) O 2 +( 2x-1 ) MgO+ H 2 O T>873 K (8)
Fig. 8.

XRD patterns of the product prepared by artificially synthesized MgTi2O5 with molten NaOH under different roasting temperatures for 1 h.

Considering that the M in NaMO2 is trivalence, it might be assumed that Ti3+ should be formed during the alkali fusion process. It is known that X-ray photoelectron spectroscopy (XPS) is often utilized to detect elements and confirm their corresponding valence state. Figure 9 illustrated the XPS of Ti element in the product obtained at 973 K for 1 h. The binding energies of the double peaks are 458.1 eV and 463.6 eV for Ti 2p3/2 and Ti 2p1/2, respectively. The energy position of this doublet only corresponds to the Ti4+ oxidation state, suggesting that none of Ti3+ existed in the product.

Fig. 9.

XPS spectrum of Ti in the alkali fusion slag obtain at 973 K for 1 h. (Online version in color.)

More importantly, from Fig. 8, it also should be noted that both diffraction peaks of MgO (2θ=43.1°) and Na2TiO3 (2θ=40.1°) decreased a lot, while the diffraction peaks of NaMO2 (2θ=16.4°, 41.1°) increased correspondingly with roasting temperature increasing from 773 K to 973 K. This phenomenon may be ascribed to that the Mg2+ ionized at relatively high temperature could partially dope into or even substitute the Ti atom of Na2TiO3 to form solid solution Na(Mg2+, Ti4+)O2, that is to say, the apparent trivalence of M in NaMO2 might be displayed by Ti4+ together with Mg2+. The results agreed well with XRD patterns (Fig. 8) and it can be inferred that higher temperature would be benefit for Mg2+ doping into/substituting the Ti atoms of Na2TiO3, leading to Ti-bearing phase transformation from Na2TiO3 to NaMO2 (M=Mg2+, Ti4+).

In order to verify the assumption, the Raman spectra were used to investigate the change of chemical bonds in the molten NaOH treatment process. Figure 10 is the Raman spectra of synthesized MgTi2O5 (Fig. 10(a)) and its corresponding alkali fusion slags obtained under different roasting temperatures (Figs. 10(b) and 10(c)). It can be clearly seen that the Raman spectrum of MgTi2O5 shows three peaks at 281, 445 and 605 cm–1, respectively. Some researchers suggested that the peak near 600 and 445 cm−1 should be ascribed to Ti–O stretching vibrations in 6-coordinated Ti4+.20,21) Meanwhile, the peaks near 281 cm−1 correspond to the bending and stretching vibration of Mg–O bonds.22) When MgTi2O5 reacted with molten NaOH at 773 K for 1 h, the peaks of the alkali fusion slag (Fig. 10(b)) didn’t show obvious change comparing to that of MgTi2O5 (Fig. 10(a)). However, the intensity of peak at 281 cm−1 decreased a lot, and in the meantime, a weak vibration peak appeared at 800 cm−1 (Fig. 10(b)). This phenomenon may be ascribed to that Ti–O band still existed as 6-coordinated Ti4+ in the obtained Na2TiO3, but the type of Mg–O bonds changed from covalent bond in MgTi2O5 to ionic bond in MgO, since MgTi2O5 can be decomposed by molten NaOH to form Na2TiO3 and MgO (Eq. (7)). Further increasing temperature to 973 K, the bending and stretching vibration peaks of Mg–O bonds at 281 cm−1 and Ti–O stretching vibration 445 and 600 cm−1 disappeared obviously, on the contrary, the intensity of vibration peak at 790 cm−1 increased dramatically (Fig. 10(c)). According to the references,20,23,24) it is known that the vibration peak at 790 cm−1 should be assigned to Ti–O2− stretching vibrations in TiO44− tetrahedral monomers. As a result of Mg2+ intercalating into Ti–O bonds, titanyl bond did not exist as 6-coordinated Ti4+, but demonstrated as Ti–O2− in TiO44− tetrahedral monomers, which combined together to form O–Ti–O or O–(Mg,Ti)–O chain units. According to the analysis of the Raman spectra, it is reasonable to conclude that MgTi2O5 can be decomposed by molten NaOH to form 6-coordinated Ti4+ and Mg2+ at 773 K. Then they combined with Na+ and O2– supplied by molten NaOH to form NaTiO3 and MgO, respectively. However, as the reaction temperature increased to 973 K, some Mg2+ would become ionized state and insert into Ti–O bonds, which made 6-coordinated Ti4+ tetrahedron become to TiO44− tetrahedral monomers and then form O–(Mg,Ti)–O chain units. During this process, the crystal structure of Na2TiO3 changed from NaCl-type crystal structure system to α-NaFeO2-type crystal structure, further confirming the transformation from Na2TiO3 to NaMO2 (M=Mg2+ and Ti4+).

Fig. 10.

Raman spectra of MgTi2O5 and alkali fusion slags obtained under different roasting temperatures for 1 h.

In addition, as for the main impurity in Ti-bearing EAF slag (e.g. MgAl2O4), the corresponding phase transformation trend was also shown in Fig. 5. It can be seen that with temperature increasing from 773 K to 973 K, the intensity of diffraction peaks of MgAl2O4 decreased obviously, suggesting higher temperature would be benefit for MgAl2O4 decomposition by molten NaOH according to the following equation:   

MgA l 2 O 4 +2NaOH=2NaAl O 2 +M g 2+ + O 2- + H 2 O (9)

Because NaOH can be used as an ionized solvent when the reaction temperature is higher than 623 K, it is reasonable to assume that the solid-liquid reaction process between MgAl2O4 and NaOH was a diffusion controlled process.18) Moreover, MgAl2O4 belongs to the cubic crystal system (unit cell parameters are a=b=c=8.09 Å and α=β=γ=90°), and it has a relative stable and dense structure since Al ions uniformly dispersed in Mg–O bonds as shown in Fig. 11(a). Therefore, increasing the amount of NaOH during the molten NaOH treatment reaction, namely, adjusting Rslag/NaOH, would lead to complete phase transformation of MgAl2O4 by molten NaOH to form NaAlO2 and MgO as illustrated in Fig. 11.

Fig. 11.

MgAl2O4 decomposed by molten NaOH to form NaAlO2 and MgO. (Online version in color.)

3.3. The Effect of Ti-bearing EAF Slag/NaOH Mass Ratio (Rslag/NaOH) on the Phase Transformation

Figure 12 gave the XRD patterns of alkaline fusion slags obtained at 973 K for 1 h when Rslag/NaOH was controlled from 1:1 to 1:1.5, respectively. It can be clearly seen that none of the diffraction peaks of anosovite solid solution or Mg2TiO4 existed in the alkali fusion slags, while some strong diffraction peaks of Na2TiO3 and NaMO2 appeared accordingly, indicating that all the Ti-bearing phases in Ti-bearing EAF slag could realize the effective phase transformation even when Rslag/NaOH was at 1:1. In addition, it is also found that the intensity of diffraction peaks of MgAl2O4 decreased dramatically when Rslag/NaOH was controlled from 1:1.2 to 1:1.5 at 973 K. This case may be ascribed to that much more amount of molten NaOH could ensure sufficient contact areas between reactants (Ti-bearing EAF slag and NaOH), leading to MgAl2O4 complete decomposition under the condition.

Fig. 12.

XRD patterns of alkaline fusion slag obtained at 973 K for 1 h with Rslag/NaOH ranging from 1:1 to 1:1.5.

More importantly, it should be noted that with Rslag/NaOH decreasing from 1:1 to 1:1.5, the intensities of diffraction peaks of Na2TiO3 (2θ=40.1°) decreased gradually while the intensities of diffraction peaks of NaMO2 (2θ=16.4°, 41.1°) increased correspondingly, suggesting that some previously formed Na2TiO3 would be spontaneously converted to NaMO2. This phenomenon may be explained as follows: when the Rslag/NaOH was adjusted from 1:1 to 1:1.5, according to Eq. (9), almost all MgAl2O4 in the slag could react with enough molten NaOH, liberating amounts of free Mg2+. In the meantime, such free Mg2+ would prefer to dope into/substitute Ti atoms of Na2TiO3 to form NaMO2 (M=Mg2+, Ti4+), rather than combine with O2– by ionic bond to produce MgO. This result agrees well with XRD patterns as shown in Fig. 12, that is to say, hardly any diffraction peaks of MgO appeared when MgAl2O4 was completely decomposed by molten NaOH. In addition, there were obvious diffraction peaks of NaAlO2 existed in XRD patterns, indicating that Al atoms in NaAlO2 could not be substituted by Mg2+ ion due to their different ionic radius (Mg2+=0.066 nm, Al3+= 0.051 nm).

Based on above analysis, in order to make the main impurity phase MgAl2O4 convert to water soluble NaAlO2 completely, and then separate from insoluble sodium titanates (Na2TiO3 and NaMO2) by water leaching, excessive amount of NaOH should be added to maintain a good reaction kinetic condition. Considering the energy consumption and material cost comprehensively, the roasting temperature and Rslag/NaOH should be controlled at 973 K and 1:1.2, respectively during the molten NaOH treatment process.

3.4. The Effect of Roasting Time on the Phase Transformation

Figure 13 illustrates XRD patterns of alkali fusion slags obtained under different time when roasting temperature was controlled at 973 K and Rslag/NaOH fixed at 1:1.2. It is shown that the diffraction peaks of all the Ti-bearing phases in Ti-bearing EAF slag (anosovite solid solution and Mg2TiO4) disappeared just within 10 min. Meanwhile, obvious diffraction peaks of corresponding sodium titanates (Na2TiO3 and NaMO2) emerged, suggesting that all the Ti-bearing phases in Ti-bearing EAF slag could be converted to sodium titanates (Na2TiO3 and NaMO2) promptly at the initial stage. Moreover, with prolonging the roasting time from 10 to 45 min, the intensities of diffraction peaks of MgAl2O4 decreased. Furthering increasing the time to 60 or 90 min, nearly all of the diffraction peaks of MgAl2O4 vanished, implying that relatively longer reaction time (>60 min) would benefit for complete decomposition of MgAl2O4.

Fig. 13.

XRD patterns of alkali fusion slag obtained at 973 K under different roasting time with Rslag/NaOH fixed at 1:1.2.

In addition, it is worth noting that the diffraction peaks of NaMO2 with weak intensity appeared within 10 min. When the roasting time increased to 30 min and further to 90 min, the intensity of diffraction peaks of NaMO2 increased gradually whereas that of Na2TiO3 decreased, and finally, even NaMO2 became the main phase in the alkali fusion slag as shown in Fig. 13. The same phenomenon also appeared when the alkali fusion process was conducted at different roasting temperatures and Rslag/NaOH. According to the above analysis, it is believed that some trivalence or divalent metal ion like Fe3+ or Mg2+ existed in Ti-bearing EAF slag could dope into/substitute the Ti atoms of Na2TiO3 to form NaMO2 (M=Ti, Fe, Mg) due to their similar ionic radius (Fe3+= 0.064 nm, Mg2+=0.066 nm and Ti4+=0.068 nm). Therefore, NaMO2 (NaFeO2, Na(Mg2+, Ti4+)O2) appeared within 10 min should be ascribed to the doped of these metal ions with low content from Ti-bearing EAF slag. With increasing of roasting time, more MgAl2O4 was decomposed to liberate more Mg2+, therefore enhanced the chance of Mg2+ doping into/substituting the Ti atoms of Na2TiO3 to form solid solution Na(Mg2+, Ti4+)O2. By this way, the amount of Na2TiO3 decreased while NaMO2 increased with roasting time pronging.

Based on the aforementioned analysis, the phase transformation mechanism during the alkali fusion process can be illustrated as Fig. 14. It is shown clearly that the main Ti-bearing phases and impurity phases in Ti-bearing EAF slag can be decomposed by NaOH to form their corresponding sodium salts. Meanwhile, some cations may dope into/substitute the Ti atoms of Na2TiO3 to form NaMO2 (M=Ti, Fe, Mg) due to their similar ionic radius. During the doping or substitution process, these cations originated from two sources. One is existing in the initial Ti-bearing EAF slag such as Fe3+ or Ti3+, which occupied a small fraction. The other (Mg2+) is from decomposition of MgTi2O5 or MgAl2O4 by NaOH with ionic state at high temperature (>873 K), which occupied high content.

Fig. 14.

The schematic diagram of phase transformation during the molten NaOH treatment process.

4. Conclusion

The present work shows that the Ti-bearing phases and the main impurity phases in Ti-bearing EAF slag can realize phase transformation effectively in molten NaOH fusion process. The phase transformation mechanism was studied systematically by thermodynamic calculation and crystal structure analysis. It is indicated that the Ti-bearing phases in Ti-bearing EAF slag can be decomposed by molten NaOH much earlier and easier than the main impurity phase MgAl2O4. The Ti-bearing phases transformed from anosovite solid solution or Mg2TiO4 to Na2TiO3 with NaCl-type crystal structure easily when the alkali fusion process was above 773 K with Ti-bearing EAF slag/NaOH mass ratio (Rslag/NaOH) below 1:1. However, the main impurity phase MgAl2O4 could not be decomposed by molten NaOH to form NaAlO2 and MgO completely until the alkali fusion reactions were conducted at 973 K for 1 h with Rslag/NaOH below 1:1.2. In addition, with increasing roasting temperature and time or decreasing Rslag/NaOH, the formed Na2TiO3 with NaCl-type crystal structure could convert to NaMO2 (M=Mg, Ti, Fe) with α-NaFeO2-type crystal structure, due to that some metal ions like Mg2+ or Fe3+ might dope into/substitute the Ti atoms of Na2TiO3 to form NaMO2 (M=Ti, Fe, Mg).

Acknowledgement

The work was financially supported by the National Natural Science Foundation of China (Nos. 51372019, 51072022 and 50874013), the National Basic Research Program of China (No. 2014CB643401).

Appendices

Table A1. Thermodynamic data.

References
 
© 2015 by The Iron and Steel Institute of Japan
feedback
Top