Next Article in Journal
Structure and Properties of High-Entropy Nitride Coatings
Next Article in Special Issue
Study of the Effect of Fluxing Ability of Flux Ores on Minimizing of Copper Losses with Slags during Copper Concentrate Smelting
Previous Article in Journal
Effect of Annealing Process on Microstructure and Magnetic Properties of FeSiBPCNbCu Nanocrystalline Soft Magnetic Powder Cores
Previous Article in Special Issue
Phase Equilibria of CaO-SiO2-La2O3-Nb2O5 System in Reducing Atmosphere
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Slag Compositions on Change Behavior of Nitrogen in Molten Steel

1
School of Metallurgy, Northeastern University, Shenyang 110819, China
2
School of Metallurgy Engineering, Liaoning Institute of Science and Technology, Benxi 117004, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(5), 846; https://doi.org/10.3390/met12050846
Submission received: 25 March 2022 / Revised: 11 May 2022 / Accepted: 13 May 2022 / Published: 16 May 2022
(This article belongs to the Special Issue Advances in Slag Metallurgy)

Abstract

:
The problem of nitrogen pickup in the smelting process of the electric arc furnace (EAF) has not been solved well. Using seven slag–steel equilibrium experiments and theoretical analysis, the relation of the foaming index and optical basicity with the nitrogen capacity of slag was clarified. Meanwhile, the effect of slag composition on the equilibrium distribution ratio of nitrogen and the mass transfer coefficient of nitrogen pickup was also studied. The results show that, with the increase in slag basicity, the nitrogen pickup amount, nitrogen pickup rate, and nitrogen equilibrium distribution ratio LN increase. Increasing the foaming index of slag and reducing its optical basicity will increase the nitrogen capacity of slag, which is conducive to hindering the nitrogen pickup of molten steel. The relationship between slag optical basicity and nitrogen capacity can be expressed as lgCN = −5.59lgΛ − 12.41. With the increase in the Al2O3 content of slag, the nitrogen pickup amount of molten steel decreases and the nitrogen pickup rate decreases. The test with MgO = 7.5% showed the highest nitrogen pickup rate and the highest nitrogen pickup mass transfer coefficient, which were 0.21 × 10−4%/min and 1.97 × 10−4 cm/s, respectively. The test with Al2O3 = 7.5% in slag showed the lowest nitrogen pickup rate and the lowest nitrogen pickup mass transfer coefficient, which were 0.08 × 10−4%/min and 1.35 × 10−4 cm/s, respectively.

1. Introduction

Electric arc furnace (EAF) is one of the main tools used for steelmaking in the world [1,2]. However, the problem of nitrogen pickup in the smelting process of EAF has not been solved well. There are several factors that affect the nitrogen content of molten steel in EAF, such as the addition of direct reduced iron, bottom blowing, foaming slag, and the mount of scrap [3,4,5,6]. Neuschütz [3] injected methane into the arc of EAF and increased the arc voltage under a constant arc current and arc length, finding that nitrogen removal was considerably accelerated after methane injection. Wei [4] reported on the effect of bottom-blowing gas on the nitrogen content of molten steel through experiments combining theoretical calculations and established kinetic models of nitrogen change in molten steel with bottom-blowing nitrogen, argon, and carbon dioxide. Derda [5] studied the effect of scrap on the nitrogen content of EAF through various experiments, finding that limiting the mass of scrap had a significant effect on reducing the nitrogen content. Brooks [6] found that injecting direct reduced iron (DRI) fines into EAF had the potential to greatly lower the nitrogen level. Fan [7] measured the dissolution of nitrogen into molten iron at 1873 K using 15N-14N isotope exchange technology and an online mass spectrometer, reporting that the nitrogen dissolution reaction apparent rate constant in pure liquid iron is ka = 4.8 × 10−6 mol∙m−2∙s∙Pa. Wang [8] explored the nitrogen absorption behavior of molten steel under different nitrogen partial pressures. The results showed that the higher the nitrogen partial pressure was, the greater the solubility of nitrogen in the molten steel was.
Some researchers have studied the behavior of nitrogen in slag. The functions of slag are to accelerate the submerged arc heating, protect the refractories from thermal radiation, and avoid the exposure of the molten steel surface to nitrogen and hydrogen [9,10]. The effect of the slag depends on the thickness of the slag layer and the foaming index [11]; the former affects the rate of nitrogen infiltration into the slag layer, while the latter affects the length of time taken by slag to isolate molten steel from the atmosphere. Meanwhile, the thickness of the slag layer is positively correlated with the foaming index and affected by the slag composition and the amount of slag. However, Matsuura [12] computed the foam height, which was limited by either the foam stability or the amount of slag. Li [13] reviewed the research progress made on denitrification slag and pointed out that adjusting the composition of slag is the key to achieving denitrification in molten steel.
Wang [14] obtained the correlative formula for nitrogen capacity, temperature, and optical basicity by regression analysis and provided two kinds of reaction mechanism for the absorption of nitrogen by slag. Their results found that slag with a high nitrogen capacity had an excellent denitrification ability. Ling [15] and Xiang [16] both studied the behavior of nitrogen in slag and molten steel through experiments; the former gave the quantitative relationship between the equilibrium distribution ratio of nitrogen and temperature, while the latter determined the mass transfer coefficient of nitrogen between slag and molten steel. Yamanaka [17] conducted a slag denitrification experiment and investigated the reaction rates. The results showed that the slag of the CaO-Al2O3-CaF2-BaF2-SiO2-MgO system was effective for use in nitrogen desorption from molten steel and that the rate of nitrogen transfer from steel to a gas phase could be increased.
Although some studies have confirmed the phenomenon of nitrogen absorption by slag, the mechanism by which slag compositions affect the change behavior of nitrogen is still unclear, and there are few related studies on this topic. Thus, the present study focused on investigating the influence of slag compositions on the nitrogen behavior in molten steel. Combining the MoSi2 furnace experiments and theoretical analysis, the relation of the foaming index and optical basicity with the nitrogen capacity of slag was clarified. Meanwhile, the effect of slag composition on the equilibrium distribution ratio of nitrogen and the mass transfer coefficient of nitrogen pickup were also studied. These results will provide guidance for slag selection and the process control of the electric arc furnace.

2. Experimental Condition and Procedure

2.1. Experimental Condition

The experimental device used was a MoSi2 furnace, as shown in Figure 1. A MgO crucible with a size of Φ53 × 65 mm was adopted as the smelting vessel, and the temperature measuring instrument used was a double platinum rhodium thermocouple (PtRH30/PtRH6S). A four-hole corundum tube with an inner diameter of 1 mm was adopted as the blow tube, and the inner diameter of the quartz tube used for sampling was 4 mm. Argon with a purity of 99.99% and nitrogen with a purity of 99.9% were used as the injection gases. The gas flow was regulated by a rotameter.
Referring to the composition of the electric arc furnace slag in a steel plant, the compositions of the slags designed for the present experiments are shown in Table 1. The raw materials used in the experiments were chemically pure reagents, which were baked in a muffle furnace at 873 K for 240 min before the experiment. The composition of the initial steel refers to the composition of the metal material fed into the EAF, and the mass fractions of the main components were as follows: C, 0.229%; Si, 0.068%; Mn, 0.160%; P, 0.010%; and S, 0.003%.

2.2. Experimental Procedure

In the experiment, 600 g of steel sample and slags with different compositions accounting for 6% of the steel mass were placed in the MgO crucible. Then, the MoSi2 furnace (Jinzhou Electric Furnace Co., Ltd., Jinzhou, China) was energized, and argon with a flow rate of 0.1 NL/min was injected into the furnace through the top blow tube. When the furnace temperature was stable at 1873 K, nitrogen was injected into the furnace at a flow rate 0.4 NL/min. Steel samples were taken with a Φ4 mm quartz tube at the time points of 0, 2, 5, 10, 15, 20, 30, 40, 50, 60, 80, 100, and 120 min. After the samples were cut and polished, the nitrogen content of the samples was detected by a LECO-TC500 Nitrogen-Oxygen Analyzer (LECO Co., St. Joseph, MI, USA). The ZSX Primus Ⅱ X-ray fluorescence spectrometer (Rigaku Co., Ltd., Akishima-shi, Japan) and Ultima Ⅳ X-ray diffractometer (XRD, Rigaku Co., Ltd., Akishima-shi, Japan) were used to analyze the slag samples collected from the crucible.

3. Experimental Results

3.1. Effect of Slag Composition on Nitrogen Content in Steel

Figure 2 and Figure 3 present the variation in the nitrogen content and nitrogen pickup rate in molten steel over time, respectively. It can be seen that within 0–5 min, the nitrogen content in molten steel increased the fastest, and the amount of nitrogen pickup was higher than 2.5 × 10−6. The nitrogen pickup rate was relatively high over 0–5 min. Except for the test of Al2O3 = 7.5%, the nitrogen pickup rate of the other tests was higher than 1.0 × 10−4%/min. The amount of nitrogen pickup and the nitrogen pickup rate were maintained at a higher level over 5–10 min. However, after 10 min, the nitrogen pickup rate of each test decreased significantly. The average nitrogen pickup rate of each test was lower than 0.21 × 10−4%/min, and the No. 4# was reduced to 0.08 × 10−4%/min. Although the nitrogen content of molten steel still showed an upward trend and the nitrogen pickup reaction in the furnace had not yet reached equilibrium at 120 min, the nitrogen pickup rate became low. Over 0–120 min, the test with the largest amount of nitrogen increased by 25.2 × 10−6.
The amount of nitrogen pickup and the nitrogen pickup rate increased with the increase in the basicity of the slag. The test of R = 2.5 had the largest pickup amount of nitrogen and the fastest nitrogen pickup rate of 0.18 × 10−4%/min. With the increase in the Al2O3 content in slag, the pickup amount of nitrogen and the nitrogen pickup rate decreased correspondingly at the same time. The test with Al2O3 = 7.5% had the lowest nitrogen pickup rate of 0.08 × 10−4%/min. The pickup amount of nitrogen and the nitrogen pickup rate increased with the increase in the MgO content in slag. The test of MgO = 7.5% had the fastest nitrogen pickup rate of 0.21 × 10−4%/min.
In the electric arc furnace, the smelting period was less than 60 min. Due to the ionization of nitrogen around the arc zone, the nitrogen pickup rate increased. The amount of nitrogen pickup was less than 18 × 10−6 in all tests within 0–60 min. Thus, if the experimental slag can achieve a submerged arc better and the average nitrogen pickup rate is less than 0.15 × 10−4%/min, it may meet the control requirements for the nitrogen content of molten steel in the electric arc furnace smelting process.

3.2. Compositions of Experimental Slags and Experimental Steels

Table 2 shows the change in the slag composition after the experiment. It was found that after the experiment, the mass fraction of FeO and MnO in slag decreased significantly, while the mass fraction of SiO2 increased. Similarly, the basicity of slag decreased to a certain extent. The chemical reactions that took place between the silicon in the molten bath and the FeO and MnO in the slag could account for the above phenomena, as shown in Equations (1) and (2). These are consistent with the reaction phenomenon in the electric arc furnace. The mass fractions of the main elements in steel after the experiment were as follows: C, 0.051–0.069%; Si, 0.005–0.009%; Mn, 0.140–0.157%; P, 0.005–0.008%; S, 0.003%.
2 ( FeO ) + [ Si ] = ( Si O 2 ) + 2 [ Fe ]
2 ( MnO ) + [ Si ] = ( Si O 2 ) + 2 [ Mn ]

3.3. The XRD Results of the Experimental Slags

The XRD results of the slag after the experiment are shown in Figure 4, from which we can see that CaMgSiO4 (monticellite), Ca3MgSi2O8 (merwinite), Ca2Al2SiO7 (calcium aluminum pyrite), Ca2SiO4 (dicalcium silicate), Al2O3, and SiO2 are in the slag.
It can be seen from Figure 4 that the slag with a basicity of 1.5 after smelting mainly contains CaMgSiO4 (monticellite), Ca3MgSi2O8 (merwinite), and Ca2Al2SiO7 (calcium aluminum pyrite). The SiO2 content is relatively large, and it cannot be completely combined with CaO; thus, some free SiO2 with a high activity will react with MgO to form monticellite and merwinite with a low melting point, as shown in Equations (3) and (4) [18]. When the basicity rises to 2.0, the amount of CaO in the slag increases and the saturated solubility of MgO decreases. A small amount of Ca2SiO4 (dicalcium silicate) is generated by the reaction between CaO and merwinite, as shown in Equation (5) [18]. Therefore, the main components are merwinite and a small amount of dicalcium silicate, and the melting point of dicalcium silicate can be as high as 2403 K. When the basicity is 2.5, the main component is dicalcium silicate with a high melting point, which is produced by the reaction of CaO and merwinite.
At 1873 K, with the increase in the Al2O3 content in the slag, the dicalcium silicate precipitated from the slag is gradually transformed into calcium aluminum pyrite, and the amount of low-melting-point merwinite is also gradually increased, as shown in Equation (6). This indicates that in the smelting process, increasing the amount of Al2O3 in slag can reduce the melting temperature of the slag (the calculation result obtained by Factsage is shown in Figure 5).
When the amount of MgO in the slag increases from 2.5% to 7.5%, the low-melting-point Al2O3 in the slag begins to transform into calcium aluminum feldspar, gradually precipitating high-melting-point SiO2 and wollastonite. Therefore, it can be seen that the melting temperature of the slag increases with the increase in MgO, but the increase in the melting point is not conducive to the covering the molten steel with slag, resulting in the greater nitrogen pickup of the molten steel, which is consistent with the research results of Meng et al. [19].
( Ca 2 + + O 2 ) ( s ) + ( Mg 2 + + O 2 ) ( s ) + SiO 2 ( s ) = CaMgSiO 4 ( s ) Δ G θ = 124766.6 + 3.768 T
3 ( Ca 2 + + O 2 ) ( s ) + ( Mg 2 + + O 2 ) ( s ) + 2 SiO 2 ( s ) = Ca 3 MgSi 2 O 8 ( s ) Δ G θ = 315469 + 24.786 T
2 ( Ca 2 + + O 2 ) + ( SiO 2 ) = ( Ca 2 SiO 4 ) Δ G θ = 160431 + 4.106 T
2 ( Ca 2 + + O 2 ) + ( Al 2 O 3 ) + ( SiO 2 ) = ( Ca 2 Al 2 SiO 7 ) ,   Δ G θ = 61964.64 60.29 T

4. Discussion

4.1. Relation of Foaming Index of Slag with Nitrogen Capacity

The research shows that there are two ways for nitrogen in the gas phase to enter the slag [20,21]:
(1) When the slag does not contain network oxides such as SiO2 and Al2O3, nitrogen enters the slag in the form of free nitrogen. The corresponding nitrogen dissolution reaction is shown in Equation (7), and the calculation formula for the nitrogen capacity of slag is shown in Equation (8).
3 2 O 2 + 1 2 N 2 = N 3 + 3 4 O 2
C N = ( N 3 ) P O 2 3 4 P N 2 1 2 = k a O 2 3 2 f N 3
(2) When the slag contains SiO2, Al2O3, and other network oxides, nitrogen reacts with SiO2 and Al2O3 network oxides in the slag and enters the slag in the form of combined nitrogen. The corresponding nitrogen dissolution reaction is shown in Equation (9), and the calculation formula for the nitrogen capacity of slag is shown in Equation (10).
3 2 O 0 + 1 2 N 2 = N 0 + 3 4 O 2
C N = ( N 0 ) P O 2 3 4 P N 2 1 2 = k a O 0 3 2 f N 0
where CN is the nitrogen capacity of slag; (N0) is the nitrogen content in slag, %; P N 2 is the partial pressure of nitrogen in the gas phase, kPa; P O 2 is the partial pressure of oxygen in the gas phase, kPa; k is the equilibrium constant of nitrogen dissolution reaction; a O 0 is the activity of O0 in slag; f N 0 is the activity coefficient of N atom in slag.
Equation (9) can be changed into Equations (11) and (13) [15] in the present slag:
3 4 SiO 2 + 1 2 N 2 = Si 0.75 N + 3 4 O 2
lg k 11 = 21926 / T + 0.737
AlO 1.5 + 1 2 N 2 = AlN + 3 4 O 2
lg k 13 = 27014 / T + 2.505
The seven slags contain network oxides such as SiO2 and Al2O3. Therefore, the thermodynamic mechanism of nitrogen reaction between gas and slag was analyzed in the second way. Since there was about five times the amount of SiO2 in the seven slags as Al2O3, k11 was selected as the equilibrium constant of nitrogen in Equation (9) at the gas–slag interface and k11 was 1.07 × 10−11 at 1873 K. The activity coefficients f N 0 and f O 0 were both 1 in the present study. The oxygen content was obtained from the slag compositions after the experiment. Thus, the nitrogen capacity of the seven slags can be calculated, as shown in Table 3.
The slag foaming index refers to the average moving time of gas in the slag foam. When the crucible diameter is greater than 30 mm, the foaming index has nothing to do with the size of the crucible and is only related to the physical properties of the slag. The larger the foaming index is, the better the foaming property of the slag is and the greater the thickness of slag. According to the study of CaO-SiO2-MgO-A12O3-FeO slag conducted by Fruehan R J et al. [22], the relationship between the foaming index and slag viscosity, surface tension, and density is shown in Equation (15):
Σ = 359 μ ρ g σ
where Σ is the slag foaming index, s; μ is the viscosity, Pa·s; ρ is the density, kg/m3; and σ is the surface tension, N/m.
The viscosity of molten slag after the experiment was calculated using the Factsage software; the calculation results are shown in Table 4.
According to reference [23], the slag density at 1673 K can be calculated by empirical Equation (16).
1 / ρ = 0.45 ( SiO 2 ) + 0.286 ( CaO ) + 0.204 ( FeO ) + 0.35 ( Fe 2 O 3 ) + 0.237 ( MnO ) + 0.367 ( MgO ) + 0.48 ( P 2 O 5 ) + 0.402 ( Al 2 O 3 )
where ρ: slag density, 103 kg/m3; (MxOy): content of oxide MxOy, %.
When the temperature T > 1673 K, the slag density at any temperature can be calculated using Equation (17).
ρ T = ρ 1673   K + 0.07 ( 1673 T 100 )
From the slag composition after the experiment, using Equations (16) and (17), the slag density under the experimental conditions can be obtained. The calculation results are shown in Table 4.
According to the chemical potential and surface energy, based on ion and molecule coexistence theory, the surface tension of slag can be calculated using Equation (18) [24,25]. For the CaO-SiO2-Al2O3-MgO-MnO-FeO slag system in the present study, Equation (18) can be expressed as Equation (21). Moreover, N i Bulk can be calculated from the mole fraction of the slag composition and the chemical equilibrium of the composite molecules based on the coexistence theory of slag structure. σ i Pure and A i are known values from references [26,27]. Combining Equations (20) and (21), N i Surf and σ can be calculated. The surface tensions of different slags are shown in Table 4.
σ = σ i Pure + R T A i ln N i Surf N i Bulk
A i = N 0 1 / 3 V i 2 / 3
N CaO Surf + N SiO 2 Surf + N Al 2 O 3 Surf + N MgO Surf + N MnO Surf + N FeO Surf = 1
σ = σ CaO Pure + R T A CaO ln N CaO Surf N CaO Bulk = σ SiO 2 Pure + R T A SiO 2 ln N SiO 2 Surf N SiO 2 Bulk = σ Al 2 O 3 Pure + R T A Al 2 O 3 ln N Al 2 O 3 Surf N Al 2 O 3 Bulk = σ MgO Pure + R T A MgO ln N MgO Surf N MgO Bulk = σ MnO Pure + R T A MnO ln N MnO Surf N MnO Bulk = σ FeO Pure + R T A FeO ln N FeO Surf N FeO Bulk
where i is the composition of slag; σ is the surface tension of slag, 10−3 N·m−1; σ i Pure is the surface tension of pure i, 10−3 N·m−1; Ai is the molar surface area of molten pure i, 10−4 m2/mol; N0 is the Avogadro constant, mol−1; Vi is the molar volume of pure molten i, L/mol; N i P is the mole fraction of i in P phase (P = surf is the surface phase, P = bulk is the bulk phase), %.
The foaming index can be obtained by substituting the viscosity, surface tension, and density of slag in Table 4 into Equation (15). Furthermore, the relationship between the nitrogen capacity of the slags and the foaming index can be obtained, as shown in Figure 6. It can be seen that the increase in the slag foaming index and the increase in the nitrogen capacity of slag hinder the nitrogen absorption of molten steel, which is consistent with the production practice results of the electric arc furnace [28]. When the slag in the electric arc furnace is foaming well, a good submerged arc can be realized, which can significantly reduce the nitrogen absorption in the smelting process.
It can be seen from Figure 6 that the foaming index of 1# slag is the largest. Slag with a good foamability is more conducive to hindering the nitrogen pickup of molten steel. This confirms the experimental results obtained for the minimum nitrogen pickup in molten steel with a slag basicity of 1.5—i.e., 1# slag—over 0–40 min, as shown in Figure 2. The foaming index of 2# slag is smaller than that of 1# slag. Since the 2# slag produced Ca2SiO4 with a high melting point (as shown in Figure 4b), it is likely to exist as solid particles at 1873 K in the slag. According to the previous research [29], a small particle size could induce the formation of foam slag, which is beneficial to hindering the nitrogen pickup of molten steel; thus, the final nitrogen mass fraction of 2# slag is the smallest (as shown in Figure 2).

4.2. Relation of Optical Basicity of Slag with Nitrogen Capacity

In 1978, Duffy J A et al. [30] proposed the concept of optical basicity (Λ). Its calculation equation is shown in Equation (22).
Λ = i = 1 n x i Λ i
According to the compositions of the experimental slags, Equation (22) can be simplified to Equation (23).
Λ = 1 B ( N C a O Λ CaO + N MgO Λ MgO + 2 N SiO 2 Λ SiO 2 + 3 N Al 2 O 3 Λ Al 2 O 3 + N FeO Λ FeO + N MnO Λ MnO )
B = N CaO + N MgO + 2 N SiO 2 + 3 N Al 2 O 3 + N FeO + N MnO
where Λ is the optical basicity of slag and Λi is the optical basicity of oxide. We set the optical basicity of pure CaO as 1; the optical basicity of other oxides is shown in Table 5 [31]; xi is the mole fraction of oxygen ions in the oxide and N is the mole fraction of each component in the slag.
Substituting the compositions of the experimental slags into Equation (23), their optical basicity can be calculated as shown in Table 6. Then, Figure 7 can be obtained. It can be seen that the nitrogen capacity decreased with the increase in the optical basicity of the slag. This phenomenon is consistent with previous studies [14,21] and can be explained by assuming that the nitride ions are combined with a network of SiO2 or Al2O3. The relationship between lgΛ and lgCN was obtained by regression, as shown in Equation (25), where the correlation coefficient R2 = 0.929.
lg C N = 5.59 lg Λ 12.41

4.3. Effect of Slag Compositions on Equilibrium Distribution Ratio of Nitrogen

In the process of electric arc furnace smelting, the nitrogen in the gas phase will react with liquid steel, as shown in Equation (26). The equilibrium distribution ratio of nitrogen between slag and steel LN is generally used as a parameter for the measurement of the denitrification ability of molten slag. The greater LN is, the stronger the denitrification ability of the slag is. The calculation of LN is shown in Equation (28).
1 2 N 2 = [ N ]   Δ G θ = 3600 + 23.9 T
k 26 = [ % N ] f N P N 2 1 / 2
L N = ( % N ) [ % N ] = f N k 26 C N P O 2 3 / 4
In accordance with the Gibbs free energy, as shown in Equation (26), the equilibrium constant k26 = 0.0448 can be obtained at 1873 K. The oxygen partial pressure at the steel–slag interface P O 2 can be calculated by Equation (30), where a [ s i ] and a ( s i O 2 ) can be calculated by Factsage, and the reaction equilibrium constant k29 can be calculated using the Gibbs free energy. The values of CN in different slags are shown in Table 3. In summary, the calculation results were substituted into Equation (28) to obtain LN. The influences of the slag compositions on LN are shown in Figure 8. It was found that LN decreases with the increase in basicity and the content of MgO in slag—that is, the denitrification ability decreases. LN increases with the increase in Al2O3—that is, the denitrification ability increases.
[ Si ] + O 2 = ( SiO 2 ) ,   Δ G θ = 947700 + 198.7 T
k 29 = a ( SiO 2 ) a [ Si ] P O 2

4.4. Effect of Slag Compositions on the Mass Transfer Coefficient of Nitrogen Pickup

The reaction speed of slag and molten steel at high temperature is fast, so it is generally considered that the mass transfer of the slag layer and the liquid phase of steel are the rate-determining steps in the nitrogen absorption reaction [32]. The nitrogen pickup reaction speed between slag and molten steel is controlled by two factors [16]: one is the mass transfer of nitrogen in slag and the other is the mass transfer of nitrogen in molten steel. The mass transfer process is expressed by mass flow J. The calculation of the nitrogen pickup mass flow of slag is shown in Equation (31).
J N s = β s ( C s C s i )
The calculation of the nitrogen pickup mass flow in molten steel is shown in Equation (32).
J N m = β m ( C m C m i )
where βs and βm are the mass transfer coefficients of nitrogen in slag and steel, cm/s; Cs and Cm are the nitrogen concentrations in slag and steel, respectively, g/cm3; and C s i and C m i are the nitrogen concentrations on the slag side and the steel side at the slag–steel interface, g/cm3, respectively.
In the steady state, the two mass flows should be equal, as shown in Equation (33).
J N s = J N m
Meanwhile, there is an equilibrium distribution of nitrogen concentration at the interface between the slag and molten steel, as shown in Equation (34).
h = C s i C m i
where h is the distribution ratio of nitrogen expressed in mass concentration (g/cm3). Then, Equation (35) can be obtained.
J N = 1 1 β s h + 1 β m ( C m C s h ) = 1 β ( C m C s h )
When expressed by the nitrogen content in molten steel, Equation (35) can be transformed into Equation (36) [17].
d [ % N ] d t = k N A V ( [ % N ] ( % N ) L N ) = k N A V ( ( 1 + W m L N W s ) [ % N ] W m L N W s [ % N ] 0 )
where kN is the total mass transfer coefficient, k N = 1 β s h + 1 β m , cm/s; Wm is the mass of molten steel, g; Ws is the mass of slag, g; A is the slag–steel interface area, cm2; and V is the volume of molten steel, cm3.
Equation (37) can be obtained by integrating it with Equation (36).
1 1 + W m L N W s ln [ % N ] 0 ( 1 + W m L N W s ) [ % N ] W m L N W s [ % N ] 0 = k N A V t
The interface area between slag and molten steel A under this experimental condition is only the surface area A1 of the molten steel in the crucible. The diameter of the crucible is 53 mm and the A/V value is 0.2321 cm−1. The mass of molten steel in the crucible is about 600 g and the mass of slag liquid is 36 g. Referring to the kinetic model shown in Equation (37), we substituted the experimental data into the model for calculation. The fit between the kinetic model and the experimental results is shown in Figure 9. It can be seen that the model calculation results fit well with the experimental data results.
The nitrogen pickup mass transfer coefficient of 2# slag with a basicity of 2.0 is 1.48 × 10−4 cm/s, meaning that the slag has a stronger ability to hinder the nitrogen pickup of molten steel. With the increase in the Al2O3 content in the slag, the viscosity of the slag increases, while the density and surface tension decrease, leading to an increase in the foaming index of the slag. Slag 4# with Al2O3 = 7.5% has the lowest nitrogen pickup mass transfer coefficient of 1.35 × 10−4 cm/s. However, with the increase in the MgO content in the slag, high-melting-point SiO2 and wollastonite are precipitated, which is not conductive to the slag covering the molten steel. Slag 6# with MgO = 7.5% has the highest nitrogen pickup mass transfer coefficient of 1.97 × 10−4 cm/s.

5. Conclusions

In the present work, the effect of EAF slag compositions on the nitrogen content of steel was evaluated. The results obtained showed that:
(1)
With the increase in the slag basicity and the MgO content in slag, the nitrogen pickup amount and the nitrogen pickup rate of molten steel increase. The test with MgO = 7.5% had the highest nitrogen pickup rate and the highest nitrogen pickup mass transfer coefficient, which were 0.21 × 10−4%/min and 1.97 × 10−4 cm/s, respectively. With the increase in the Al2O3 content in slag, the nitrogen pickup amount of molten steel decreased and the nitrogen pickup rate decreased. The test with Al2O3 = 7.5% in slag has the lowest nitrogen pickup rate and the lowest nitrogen pickup mass transfer coefficient of 0.08 × 10−4%/min and 1.35 × 10−4 cm/s, respectively.
(2)
Increasing the foaming index of slag and reducing the optical basicity of slag will increase the nitrogen capacity of slag, which is conducive to hindering the nitrogen pickup of molten steel. The relationship between the slag optical basicity and nitrogen capacity can be expressed as: lgCN = −5.59lgΛ − 12.41.
(3)
The nitrogen equilibrium distribution ratio LN between slag and molten steel decreases with the increase in the basicity, increases with the increase in Al2O3, and decreases with the increase in MgO.

Author Contributions

Data curation, D.Z. and J.W.; Funding acquisition, D.Z. and H.Z.; Investigation, J.W. and D.Z.; Validation, J.W., D.Z. and L.H.; Writing—original draft, D.Z. and J.W.; Writing—review and editing, D.Z. and L.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (grant number 51734003), and Support Plan for Innovative Talents in Colleges and Universities of Liaoning Provincial Department of Education (grant number LR2020066).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhu, R.; Wu, X.T.; Wei, G.S.; Tian, B.H. Development of green and intelligent technologies in electric arc furnace steelmaking processes. Iron Steel 2019, 54, 9–20. (In Chinese) [Google Scholar]
  2. Jiang, Z.H.; Yao, C.L.; Zhu, H.C.; Pan, T. Technology development trend in electric arc furnace steelmaking. Iron Steel 2020, 55, 1–12. (In Chinese) [Google Scholar]
  3. Neuschütz, D.; Spirin, D. Nitrogen removal and arc voltage increase in EAF steelmaking by methane injection into the arc. Steel Res. Int. 2003, 74, 19–25. [Google Scholar] [CrossRef]
  4. Wei, G.S.; Zhu, R.; Dong, K.; Li, Z.Z.; Yang, L.Z.; Wu, X.T. Influence of bottom-blowing gas species on the nitrogen content in molten steel during the EAF steelmaking process. Ironmak. Steelmak. 2018, 45, 839–846. [Google Scholar] [CrossRef]
  5. Derda, W.; Siwka, J.; Nowosielski, C.Z. Controlling of the nitrogen content during EAF–technology and continuous casting of steel. Arch. Metall. Mater. 2008, 53, 523–529. [Google Scholar]
  6. Brooks, G.; Irons, G.; Anghelina, D. New approach to nitrogen control in EAF steelmaking. In Proceedings of the 4th High Temperature Processing Symposium, Melbourne, Australia, 6–7 February 2012. [Google Scholar]
  7. Fan, Y.W.; Hu, X.J.; Wang, P.D.; Li, Y. Effects of carbon, aluminum and silicon on the dissolution rate of nitrogen into molten iron. Chin. J. Eng. 2020, 42, 34–38. (In Chinese) [Google Scholar]
  8. Wang, J.X.; Zhan, D.P.; Liu, X.; Huang, Q.Q.; Jiang, Z.H.; Zhang, H.S. Research on nitrogen absorption behavior of molten steel during submerged nitro-blowing process in EAF. J. Mater. Metall. 2020, 19, 265–269. (In Chinese) [Google Scholar]
  9. Luz, A.P.; Tomba Martinez, A.G.; López, F.; Bonadia, P.; Pandolfelli, V.C. Slag foaming practice in the steelmaking process. Ceram. Int. 2018, 44, 8727–8741. [Google Scholar] [CrossRef]
  10. Zhang, B.Y.; Tian, B.H.; Wei, G.S. New process of nitrogen content change and control technology in EAF steelmaking process. Ind. Heat. 2020, 49, 19–24. (In Chinese) [Google Scholar]
  11. Le, K.X.; Dong, Y.C.; Wang, S.J.; Sun, W.; Mao, Y. Experimental study of foaming slag. J. Baotou Univ. Iron Steel Technol. 1999, 18, 256–259. (In Chinese) [Google Scholar]
  12. Matsuura, H.; Fruehan, R.J. Slag foaming in an electric arc furnace. ISIJ Int. 2009, 49, 1530–1535. [Google Scholar] [CrossRef] [Green Version]
  13. Li, Q.Q.; Wu, H.J.; Wang, Z. Research progress on denitrification slag. Foundry Technol. 2018, 39, 2620–2624. (In Chinese) [Google Scholar]
  14. Wang, M.; Bao, Y.P.; Cui, H.; Wu, H.J.; Liu, J.H. Nitrogen behavior in CaO-SiO2-Al2O3-MgO refining slags. J. Univ. Sci. Technol. Beijing 2010, 32, 175–178. (In Chinese) [Google Scholar]
  15. Ling, T.Y.; Xu, K.D.; Jiang, G.C. Nitrogen stability in gas-slag and slag-metal systems. J. Shanghai Univ. Technol. 1991, 12, 162–169. (In Chinese) [Google Scholar]
  16. Xiang, C.X.; Gammal, T.E. Measurement of transport coefficients of nitrogen between slag and liquid steel. J. Univ. Sci. Technol. Beijing 1997, 19, 282–286. (In Chinese) [Google Scholar]
  17. Yamanaka, R.; Ogawa, K.; Iritani, H.; Koyama, S. Denitrogenization mechanism from molten steel by flux treatment. ISIJ Int. 1992, 32, 136–141. [Google Scholar] [CrossRef] [Green Version]
  18. Shen, F.M. Physical Chemistry of Metallurgy, 1st ed.; Higher Education Press: Beijing, China, 2017; pp. 1–292. [Google Scholar]
  19. Meng, Q.Y.; Hong, X.; Qiu, X.L.; Zhang, M.M.; Li, Q. Low fluoride foamed slag for manufacturing stainless steel in EAF. J. Iron Steel Res. 2007, 19, 16–20. (In Chinese) [Google Scholar]
  20. Min, D.J.; Fruehan, R.J. Nitrogen solution in BaO-B2O3 and CaO-B2O3 slags. Metall. Mater. Trans. B 1990, 21, 1025–1032. [Google Scholar] [CrossRef]
  21. Ito, K.; Fruehan, R.J. Thermodynamics of nitrogen in CaO-SiO2-Al2O3 slags and its reaction with Fe-Csat. Metall. Mater. Trans. B 1988, 19, 419–425. [Google Scholar] [CrossRef]
  22. Jiang, R.; Fruehan, R.J. Slag foaming in bath smelting. Metall. Mater. Trans. B 1991, 22, 481–489. [Google Scholar] [CrossRef]
  23. Xin, J.J.; Gan, L.; Jiao, L.; Lai, C.B. Accurate density calculation for molten slags in SiO2-Al2O3-CaO-MgO systems. ISIJ Int. 2017, 57, 1340–1349. [Google Scholar] [CrossRef] [Green Version]
  24. Cheng, G.G.; Liao, N.B. Calculation model for surface tension of slag melt. J. Iron Steel Res. Int. 1999, 6, 17–20. [Google Scholar]
  25. Bulter, J.A.V. The thermodynamics of the surfaces of solutions. Proc. R. Soc. Lond. A 1932, 135, 348–375. [Google Scholar]
  26. Nakamoto, M.; Kiyose, A.; Tanaka, T.; Holappa, L.; Hämäläinen, M. Evaluation of the surface tension of ternary silicate melts containing Al2O3, CaO, FeO, MgO or MnO. ISIJ Int. 2007, 47, 38–43. [Google Scholar] [CrossRef] [Green Version]
  27. Hanao, M.; Tanaka, T.; Kawamoto, M.; Takatani, K. Evaluation of surface tension of molten slag in multi-component systems. ISIJ Int. 2007, 47, 935–939. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, X.Y.; Song, Z.Y. Exploration of manufacture progress for reducing line pipe nitrogen content in electric furnace. Ind. Heat. 2019, 48, 14–16. (In Chinese) [Google Scholar]
  29. Shahbazian, F.; Du, S.C.; Seetharaman, S. The effect of addition of Al2O3 on the viscosity of CaO-“FeO”-SiO2-CaF2 slags. ISIJ Int. 2002, 42, 155–162. [Google Scholar] [CrossRef] [Green Version]
  30. Duffy, J.A.; Ingram, M.D.; Sommerville, I.D. Acid-base properties of molten oxides and metallurgical slags. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1978, 74, 1410–1419. [Google Scholar] [CrossRef]
  31. Ling, T.Y.; Xu, K.D.; Jiang, G.C. Thermodynamics of nitrogen in CaO-SiO2-Al2O3. J. Iron Steel Res. 1990, 2, 13–19. (In Chinese) [Google Scholar]
  32. Chang, H.M.; Meng, Q.Y.; Fu, J. Removal of nitrogen from liquid metal. J. Iron Steel Res. 1995, 7, 73–79. (In Chinese) [Google Scholar]
Figure 1. MoSi2 furnace device.
Figure 1. MoSi2 furnace device.
Metals 12 00846 g001
Figure 2. Results of the nitrogen pickup of molten steel in different tests (a) different basicity slags; (b) different Al2O3 content slags; (c) different MgO content slags.
Figure 2. Results of the nitrogen pickup of molten steel in different tests (a) different basicity slags; (b) different Al2O3 content slags; (c) different MgO content slags.
Metals 12 00846 g002
Figure 3. Nitrogen pickup rate of molten steel in different tests (a) different basicity slags; (b) different Al2O3 content slags; (c) different MgO content slags.
Figure 3. Nitrogen pickup rate of molten steel in different tests (a) different basicity slags; (b) different Al2O3 content slags; (c) different MgO content slags.
Metals 12 00846 g003
Figure 4. XRD results of the slags after smelting with different tests: (a) 1#, R = 1.5; (b) 2#, R = 2.0; (c) 3#, R = 2.5; (d) 4#, Al2O3 = 7.5%; (e) 5#, Al2O3 = 2.5%; (f) 6#, MgO = 7.5%; (g) 7#, MgO = 2.5%.
Figure 4. XRD results of the slags after smelting with different tests: (a) 1#, R = 1.5; (b) 2#, R = 2.0; (c) 3#, R = 2.5; (d) 4#, Al2O3 = 7.5%; (e) 5#, Al2O3 = 2.5%; (f) 6#, MgO = 7.5%; (g) 7#, MgO = 2.5%.
Metals 12 00846 g004
Figure 5. Phase diagram of CaO-SiO2-Al2O3-5%MgO.
Figure 5. Phase diagram of CaO-SiO2-Al2O3-5%MgO.
Metals 12 00846 g005
Figure 6. Relationship between nitrogen capacity and foaming index.
Figure 6. Relationship between nitrogen capacity and foaming index.
Metals 12 00846 g006
Figure 7. Optical basicity and nitrogen capacity.
Figure 7. Optical basicity and nitrogen capacity.
Metals 12 00846 g007
Figure 8. Equilibrium distribution ratio of nitrogen.
Figure 8. Equilibrium distribution ratio of nitrogen.
Metals 12 00846 g008
Figure 9. Fitting of the kinetic model and experimental results (a) Different basicity slags; (b) Different Al2O3 content slags; (c) Different MgO content slags.
Figure 9. Fitting of the kinetic model and experimental results (a) Different basicity slags; (b) Different Al2O3 content slags; (c) Different MgO content slags.
Metals 12 00846 g009
Table 1. Compositions of designed slag, mass%.
Table 1. Compositions of designed slag, mass%.
No.CaOSiO2Al2O3MgOMnOFeOR
1#42.0%28.0%5.0%5.0%7.0%13.0%1.5
2#46.5%23.5%5.0%5.0%7.0%13.0%2.0
3#50.0%20.0%5.0%5.0%7.0%13.0%2.5
4#45.0%22.5%7.5%5.0%7.0%13.0%2.0
5#48.5%24.0%2.5%5.0%7.0%13.0%2.0
6#45.0%22.5%5.0%7.5%7.0%13.0%2.0
7#48.5%24.0%5.0%2.5%7.0%13.0%2.0
Table 2. Slag compositions in different tests after smelting, mass%.
Table 2. Slag compositions in different tests after smelting, mass%.
No.CaOSiO2Al2O3MgOMnOFeOR
1#41.8%32.2%7.1%11.7%4.4%2.8%1.30
2#51.6%31.1%4.6%6.1%3.8%2.8%1.66
3#55.4%27.6%7.5%3.8%2.9%2.8%2.01
4#48.0%29.6%8.7%7.6%4.0%2.1%1.62
5#54.9%31.9%3.1%4.5%3.2%2.4%1.72
6#52.2%31.8%4.6%6.5%2.9%2.0%1.64
7#53.9%31.7%4.6%5.6%2.6%1.6%1.70
Table 3. Nitrogen capacity of slags.
Table 3. Nitrogen capacity of slags.
No.1#2#3#4#5#6#7#
CN, 10−121.851.591.391.661.491.531.62
Table 4. Physical parameters of the slags.
Table 4. Physical parameters of the slags.
No.1#2#3#4#5#6#7#
μ, Pa∙s (1873 K)0.0500.0420.0370.0440.0390.0410.042
ρ, 103 kg∙m−32.9122.9823.0392.9703.0022.9782.994
σ, 10−3 N∙m−1528.386528.393528.406527.944528.804529.081527.666
Table 5. Optical basicity of oxides.
Table 5. Optical basicity of oxides.
OxideCaOMgOSiO2Al2O3MnOFeO
Λ10.920.480.680.950.93
Table 6. Optical basicity of the experimental slag.
Table 6. Optical basicity of the experimental slag.
No.1#2#3#4#5#6#7#
Λ0.7590.7770.8000.7760.7820.7830.775
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhan, D.; Wang, J.; Huang, L.; Zhang, H. Effect of Slag Compositions on Change Behavior of Nitrogen in Molten Steel. Metals 2022, 12, 846. https://doi.org/10.3390/met12050846

AMA Style

Zhan D, Wang J, Huang L, Zhang H. Effect of Slag Compositions on Change Behavior of Nitrogen in Molten Steel. Metals. 2022; 12(5):846. https://doi.org/10.3390/met12050846

Chicago/Turabian Style

Zhan, Dongping, Jiaxi Wang, Luoyi Huang, and Huishu Zhang. 2022. "Effect of Slag Compositions on Change Behavior of Nitrogen in Molten Steel" Metals 12, no. 5: 846. https://doi.org/10.3390/met12050846

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