Skip to content
BY 4.0 license Open Access Published by De Gruyter September 10, 2020

Oxidizability characterization of slag system on the thermodynamic model of superalloy desulfurization

  • Wanming Li , Yulei Sun , Shaopeng Wu and Ximin Zang EMAIL logo

Abstract

The oxidizability characterization method of slag system has a decisive influence on the accuracy of the desulfurization model. By carrying out the balance experiments of melting system composed of CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO slag system and Inconel 718 superalloy, the influence of Al2O3 and FeO on desulfurization distribution ratio at 1,773 K was studied. Based on the ion and molecule coexistence theory in slag, the effects of three oxidizability characterization methods, [Fe]–[O] balance, [Al]–[O] balance, and [Fe]–[Al]–[O] balance, on the accuracy of desulfurization thermodynamic model were studied. The results show that the effect of FeO on desulfurization distribution ratio is more significant than that of Al2O3 when the FeO content in slag is greater than 0.014%. When the FeO content in the slag is less than 0.014%, the effect of Al2O3 on the slag oxidizability will be greater than that of FeO. The calculated value of [Fe]–[Al]–[O] balance model is closer to the experimental-measured value compared to [Fe]–[O] balance model and [Al]–[O] balance model alone at 1,923 K.

1 Introduction

Inconel 718 alloy is widely used in aerospace and other fields due to its excellent properties of high temperature resistance, corrosion resistance, and high strength. Sulfur is a harmful element in nickel-based superalloys, which can reduce the tensile plasticity, creep resistance, and high-temperature fatigue life of superalloys [1,2,3,4]. In addition, Bricknell et al. [5] indicated that the oxidation of sulfides on the surface of alloys caused by oxygen infiltration could destroy the dense oxide film formed by aluminum and chromium elements, resulting in a brittle substrate surface. Moreover, sulfur can degrade the hot plasticity, hot ductility, and fatigue property and increase the solidification segregation of superalloy [6,7].

The oxidizability of slag has a complex effect on the sulfur content ratio [8]. Li et al. [9] studied the influence of slag system on sulfur in the electroslag remelting process and concluded that reducing the easily oxidized elements in the slag can enhance the desulfurization rate. Geng et al. [10] proposed that unstable oxides (FeO, Cr2O3, MnO, SiO2, etc.) in slag would increase the oxidizability of slag, affecting the dissolved oxygen content in superalloy, and therefore further affecting the desulfurization reaction. Hou et al. [11] showed that slag system containing CaO would lead to enhanced water absorption in the electroslag remelting process, easily causing higher oxygen levels of superalloy and affecting slag desulfurization.

In order to explain the relationship between slag oxidizing and desulfurization ability, scholars used [Fe]–[O] balance and [Al]–[O] balance to characterize the oxygen content in the molten metal and then built a desulfurization thermodynamic model. But the model calculation was not accurate enough. Liu et al. [12] established a mass concentration model of a six-component (CaO–MgO–FeO–CaF2–SiO2–Al2O3) slag and characterized the dissolved oxygen content based on the [Fe]–[O] balance. However, the oxygen content calculated by the model is lower than the results of industrial tests, which also leads to the inaccuracy of the desulfurization thermodynamic calculation model. Taniguchi et al. [13] characterized the dissolved oxygen content during steelmaking based on [Fe]–[O] balance, but the calculation results also demonstrated a big error with the actual results. Studies have shown that the increase in oxygen content is mainly caused by the decomposition of unstable oxides in the slag. In the metallurgical process, the FeO content is usually used to characterize the slag oxidizability. However, the slag temperature in the electroslag remelting process can reach about 2,173 K, which is much higher than the traditional metallurgical process, and the slag system used in the electroslag remelting process usually contains more Al2O3 (about 30%). These are conducive to the decomposition of Al2O3 and make Al2O3 an unstable substance [14]. In addition, the required FeO content in the slag in the electroslag remelting process is very low and usually less than 0.5%. At this time, the effect of Al2O3 on the slag oxidizabilty should not be ignored. Yang et al. [15] conducted the [Al]–[O] balance experiment and concluded that the calculated dissolved oxygen content based on [Al]–[O] balance has a large error with the experimental results. The desulfurization thermodynamic model of Inconel 718 superalloy electroslag remelting process was studied through a slag–metal balance experiment. The [Fe]–[Al]–[O] balance was adopted to characterize the oxidizabilty of slag. From the thermodynamic point of view, the influence law of slag oxidizability on desulfurization is clarified, which could provide a theoretical basis for precise control of sulfur content in superalloy electroslag remelting process.

In this article, the ion and molecular coexistence theory of slag structure was used to establish the mass–concentration control equations of Inconel 718 superalloy and CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO slag system. The [Fe]–[O] balance, [Al]–[O] balance, and [Fe]–[Al]–[O] balance were used to build the slag–metal desulfurization thermodynamic model. The effect of three slag oxidizability characterization methods on the desulfurization thermodynamic model was evaluated through experiments.

2 Experiments

The schematic diagram of the experimental device is shown in Figure 1(a). A magnesia crucible with an inner diameter of 60 mm and a height of 100 mm was used in this experiment. The temperature of the MoSi2 furnace is set to 1773 K. The Inconel 718 superalloy and slag used in the experiment weight 150 and 50 g, respectively. Table 1 shows the chemical compositions of the Inconel 718 superalloy, and Table 2 shows the compositions of slag. When one component in the slag changes, the remaining components will increase or decrease at the original ratio. In the experiment, chemically pure reagents were used as the components of slag. All used reagents were dried in an oven at 1,073 K for 4 h to remove the water vapor. We refer to the experimental method in Duan’s work [16] wherein the alloy sample was added to the crucible and placed in the constant temperature zone of the MoSi2 furnace. The temperature was kept at 1,773 K for 20 min to completely melt the sample and then the stopper was pulled up to let the slag flow down, as shown in Figure 1(b) and (c). To avoid the influence of oxygen in the air during the melting process, argon was injected into the furnace with a flow rate of 2.5 L min−1 throughout the experiment. When the reaction time reached 80 min, the molten slag and superalloy samples were sucked up by the quartz tube and then the crucible was removed and cooled by water. The sulfur content in the slag was analyzed by X-ray fluorescence and determined by a carbon–sulfur analyzer. The measured sulfur content results of the slag and alloy sucked out are listed in Tables 3 and 4, respectively.

Figure 1 Schematic diagram of the experiment (a) experimental device, (b) melting stage, and (c) slag–metal reaction stage.
Figure 1

Schematic diagram of the experiment (a) experimental device, (b) melting stage, and (c) slag–metal reaction stage.

Table 1

Chemical composition of Inconel 718 superalloy (wt%)

ElementCSiMnMoNiCrNbAlTiSFe
Content0.040.240.193.0851.9618.535.670.671.110.013Bal.
Table 2

Chemical compositions of slags used in the experiment (wt%)

Slag systemCaF2CaOAl2O3SiO2TiO2MgOFeO
S049.9020.020.043.014.013.010
S149.8020.0020.003.004.003.000.20
S249.7019.9619.962.993.992.990.40
S349.5519.9019.902.983.982.980.70
S447.9319.25232.893.852.890.19
S546.6918.75252.813.752.810.18
S645.4418.25272.743.652.740.18
S744.2017.75292.663.552.660.17
S842.9517.25312.593.452.590.17
Table 3

The sulfur content in final slag samples (wt%)

SlagsS0S1S2S3S4S5S6S7S8
w(S)0.09840.04080.03610.03560.04120.03990.03920.03890.0387
Table 4

The sulfur content in final alloy samples (wt%)

Samples012345678
w[S]0.000710.000630.000720.000820.000750.000840.000950.001060.00112

3 Establishment of mass action concentration equations of CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO slags based on the ion and molecule coexistence theory

The ion and molecule coexistence theory [17,18,19,20,21] of slag structure can be used to describe the reactions between molecules and ions in the molten slags. Complex molecules in the slag are formed by the reactions of simple molecules and simple ions in dynamic balance. Assuming that the slag components follow the mass action law, the activity of each component in the slag can be expressed by a mass action concentration. Combining the currently known thermodynamic data and mass action law, the action concentration of each structural unit in the slag can be predicted more accurately.

3.1 Model of mass action concentrations of structural units and ion couples

According to the ion and molecular coexistence theory, the phase composition of slag is very important for the establishment of the desulfurization thermodynamic model. According to the phase diagrams of CaO–SiO2, CaO–Al2O3, TiO2–Al2O3, TiO2–CaO, CaO–Al2O3–SiO2, CaO–SiO2–CaF2, CaO–Al2O3–CaF2, CaO–SiO2–TiO2, CaO–FeO–SiO2, CaO–MgO–SiO2, and MgO–Al2O3–SiO2 [22], the structural units from 1,473 to 1,973 K [23] are listed in Table 5.

Table 5

The structural unit and mass action concentrations existed in the balance of the slag system CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO

Structural unitsMass action concentrationsStructural unitsMass action concentrations
Ca2+ + O2−N12FeO·SiO2Nc15
Mg2+ + O2−N2CaO·TiO2Nc16
Ca2+ + 2FN33CaO·2TiO2Nc17
Fe2+ + O2−N44CaO·3TiO2Nc18
SiO2N5MgO·TiO2Nc19
TiO2N6MgO·2TiO2Nc20
Al2O3N72MgO·TiO2Nc21
3CaO·SiO2Nc1Al2O3·TiO2Nc22
2CaO·SiO2Nc22MgO·2Al2O3·5SiO2Nc23
CaO·SiO2Nc32CaO·Al2O3·SiO2Nc24
3CaO·Al2O3Nc4CaO·Al2O3·2SiO2Nc25
12CaO·7Al2O3Nc5CaO·MgO·2SiO2Nc26
CaO·Al2O3Nc62CaO·MgO·2SiO2Nc27
CaO·2Al2O3Nc73CaO·MgO·2SiO2Nc28
CaO·6Al2O3Nc83CaO·3Al2O3·CaF2Nc29
2MgO·SiO2Nc911CaO·7Al2O3·CaF2Nc30
MgO·SiO2Nc10CaO·MgO·SiO2Nc31
MgO·Al2O3Nc113CaO·2SiO2·CaF2Nc32
3Al2O3·2SiO2Nc12CaO·SiO2·TiO2Nc33
3CaO·2SiO2Nc13CaO·FeO·SiO2Nc34
FeO·Al2O3Nc14

The molten slag was regarded as an ideal solution in dynamic balance reactions between the simple ions and the simple molecules. The weight of the initial slag system was set to be 100 g. To represent the chemical composition of slags, the variables, b1=nCaO0, b2=nMgO0, b3=nCaF20, b4=nFeO0, b5=nSiO20, and b6=nTiO20, b7=nAl2O30, were assigned to represent the number of moles of CaO, MgO, CaF2, FeO, SiO2, TiO2, and Al2O3 in the slag system CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO. The ion and molecule coexistence theory hold that each ion couple is electrically neutral and can be dissociated into separate cations and anions. Each ion pair’s balance number of moles is equal to the summation of the cations and the anions in the ion couple.

The total number of balance moles of all structural units in 100 g slag ni can be expressed as follows:

(1)ni=2n1+2n2++n7+nc1+nc2++nc30

According to the ion and molecule coexistence theory, the ratio of the balance number of moles of a structural unit i to the total number of moles of all the structural units in the balance system is the mass action concentration of the structural unit i in the molten slag, which is called Ni. It can be calculated as follows:

(2)Ni=ni/ni

When the system reaches a dynamic balance, each simple molecule and complex molecule can only provide one structural unit. Therefore, the mass action concentrations of all structure units in the form of ions, simple molecules, and complex molecules in CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO slag can be calculated from equation (2). However, mass action concentrations of ion couples, such as (Ca2+ + O2−), should be calculated as follows:

(3)NCaO=NCa2+,CaO+NO2,CaO=nCa2+,CaO+nO2,CaOni=2nCaOni

Assuming the conservation of mass and dynamic balances for the reactions of the slag structure units, the activity model of slag can be established. The ions and structural units can form complex molecules through chemical reactions in the slag system CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO. The chemical reactions, the standard molar Gibbs free energies ΔrGm,ciθ, and the mass action concentrations of possibly formed complex molecules are shown in Table 6. The standard molar Gibbs free energies for the reactions were cited from the literatures [24,25] or calculated with the FactSage software.

Table 6

Chemical reaction formulas, standard molar Gibbs free energy, and mass action concentrations of possibly formed complex molecules in the system CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO at different temperatures

ReactionsΔrGm,ciθ(Jmol1)Ni
3(Ca2+ + O2−) + (SiO2) = (3CaO·SiO2)−1,18,826 to 6.694TNc1=Kc1N13N5
2(Ca2+ + O2−) + (SiO2) = (2CaO·SiO2)−1,02,090 to 24.267TNc2=Kc2N12N5
(Ca2+ + O2−) + (SiO2) = (CaO·SiO2)−21,757 to 36.819TNc3=Kc3N1N5
3(Ca2+ + O2−) + (Al2O3) = (3CaO·Al2O3)−21,757 to 29.288TNc4=Kc4N13N7
12(Ca2+ + O2−) + 7(Al2O3) = (12CaO·7Al2O3)6,17,977–612.119TNc5=Kc5N112N77
(Ca2+ + O2−) + (Al2O3) = (CaO·Al2O3)59,413–59.413TNc6=Kc6N1N7
(Ca2+ + O2−) + 2(Al2O3) = (CaO·2Al2O3)−16,736 to 25.522TNc7=Kc7N1N72
(Ca2+ + O2−) + 6(Al2O3) = (CaO·6Al2O3)−22,594 to 31.798TNc8=Kc8N1N76
2(Mg2+ + O2−) + (SiO2) = (2MgO·SiO2)−56,902 to 3.347TNc9=Kc9N22N5
(Mg2+ + O2−) + (SiO2) = (MgO·SiO2)23,849–29.706TNc10=Kc10N2N5
(Mg2+ + O2−) + (Al2O3) = (MgO·Al2O3)−18,828 to 6.276TNc11=Kc11N2N7
3(Al2O3) + 2(SiO2) = (3Al2O3·2SiO2)−4354.27 to 10.467TNc12=Kc12N73N52
3(Ca2+ + O2−) + 2(SiO2) = (3CaO·2SiO2)−2,36,814 + 9.623TNc13=Kc13N13N52
(Fe2+ + O2−) + (Al2O3) = (FeO·Al2O3)−33272.8 + 6.1028TNc14=Kc14N7N4
2(Fe2+ + O2−) + (SiO2) = (2FeO·SiO2)−28595.84 + 3.349TNc15=Kc15N42N5
(Ca2+ + O2−) + (TiO2) = (CaO·TiO2)−79,900 to 3.35TNc16=Kc16N1N6
3(Ca2+ + O2−) + 2(TiO2) = (3CaO·2TiO2)−2,07,100 to 11.35TNc17=Kc17N13N62
4(Ca2+ + O2−) + 3(TiO2) = (4CaO·3TiO2)−2,92,880 to 17.573TNc18=Kc18N14N63
(Mg2+ + O2−) + (TiO2) = (MgO·TiO2)−26376.8 + 3.14TNc19=Kc19N2N6
(Mg2+ + O2−) + 2(TiO2) = (MgO·2TiO2)−27632.9 + 0.63TNc20=Kc20N2N62
2(Mg2+ + O2−) + (TiO2) = (2MgO·TiO2)−25539.5 + 1.26TNc21=Kc21N22N6
(Al2O3) + (TiO2) = (Al2O3·TiO2)−25,270 + 3.924TNc22=Kc22N7N6
2(Mg2+ + O2−) + 2(Al2O3) + 5(SiO2) = (2MgO·2Al2O3·5SiO2)−14,422 to 14.808TNc23=Kc23N22N72N55
2(Ca2+ + O2−) + (Al2O3) + (SiO2) = (2CaO·Al2O3·SiO2)−1,16,315 to 38.911TNc24=Kc24N12N7N5
(Ca2+ + O2−) + (Al2O3) + 2(SiO2) = (CaO·Al2O3·2SiO2)−4,184 to 73.638TNc25=Kc25N1N7N52
(Ca2+ + O2−) + (Mg2+ + O2−) + 2(SiO2) = (CaO·MgO·2SiO2)−80,387 to 51.916TNc26=Kc26N1N2N52
2(Ca2+ + O2−) + (Mg2+ + O2−) + 2(SiO2) = (2CaO·MgO·2SiO2)−73,688 to 63.639TNc27=Kc27N12N2N52
3(Ca2+ + O2−) + (Mg2+ + O2−) + 2(SiO2) = (3CaO·MgO·2SiO2)−2,05,016 to 31.798TNc28=Kc28N13N2N52
3(Ca2+ + O2−) + 3(Al2O3) + (Ca2+ + 2F) = (3CaO·3Al2O3·CaF2)−44,492 to 73.15TNc29=Kc29N13N73N3
11(Ca2+ + O2−) + 7(Al2O3) + (Ca2+ + 2F) = (11CaO·7Al2O3·CaF2)−2,28,760 to 155.8TNc30=Kc30N111N77N3
(Ca2+ + O2−) + (Mg2+ + O2−) + (SiO2) = (CaO·MgO·SiO2)−124766.6 + 3.768TNc31=Kc31N1N2N5
3(Ca2+ + O2−) + 2(SiO2) + (Ca2+ + 2F) = (3CaO·2SiO2·CaF2)−2,55,180 to 8.20TNc32=Kc32N13N52N3
(Ca2+ + O2−) + (SiO2) + (TiO2) = (CaO·SiO2·TiO2)−1,14,683 + 7.32TNc33=Kc33N1N5N6
(Ca2+ + O2−) + (Fe2+ + O2−) + (SiO2) = (CaO·FeO·SiO2)−72996.8 to 29.3169TNc34=Kc34N1N4N5

Based on the balance number of moles and mass action concentration of each structural unit of slag, the mass conservation equation is established as follows:

(4)b1=0.5N1+3Nc1+2Nc2+Nc3+3Nc4+12Nc5+Nc6+Nc7+Nc8+3Nc13+Nc16+3Nc17+4Nc18+2Nc24+Nc25+Nc26+2Nc27+3Nc28+3Nc29+11Nc30+Nc31+3Nc32+Nc33+Nc34ni=nCaO0
(5)b2=0.5N2+2Nc9+Nc10+Nc11+Nc19+Nc20+2Nc21+2Nc23+Nc26+Nc27+Nc28+Nc31ni=nMgO0
(6)b3=1/3N3+Nc29+Nc30+Nc32ni=nCaF20
(7)b4=(0.5N4+Nc14+2Nc15+Nc34)ni=nFeO0
(8)b5=N5+Nc1+Nc2+Nc3+Nc9+Nc10+2Nc12+2Nc13+Nc15+5Nc23+Nc24+2Nc25+2Nc26+2Nc27+2Nc28+Nc31+2Nc32+Nc33+Nc34ni=nSiO20
(9)b6=N6+Nc16+2Nc17+3Nc18+Nc19+2Nc20+Nc21+Nc22+Nc33ni=nTiO20
(10)b7=N7+Nc4+7Nc5+Nc6+2Nc7+6Nc8+Nc11+3Nc12+Nc14+Nc22+2Nc23+Nc24+Nc25+3Nc29+7Nc30ni=nAl2O30

Because the sum of the mole fractions of all structural units in slag is 1 under balance conditions, the following formula can be obtained:

(11)N1+N2+N3+N4+N5+N6+N7+Nc1+Nc2++Nc34=1

Equations (4)–(11) govern the calculation of the mass action concentrations of structural units or ion couples in slag system CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO. There are eight unknowns as in N1, N2, N3, N4, N5, N6, and N7 and ni with eight independent equations. Using MATLAB 7.0 to solve these higher order nonlinear equations, the activity of each component in slag was obtained.

3.2 Verification of mass action concentration model

The mass action concentration model of slags is also called the activity model of slag components. Because the measured activity values of slag CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO have rarely been reported, another common slag system studied by researchers [26,27] was used to verify the new model. The components of the common slag system were brought into the model and then the calculated activity values were compared with previous experimental results, as seen in Figure 2.

Figure 2 Comparison between calculated values of the activity model and the tested values. (a) FeO, (b) CaO, and (c) Al2O3.
Figure 2

Comparison between calculated values of the activity model and the tested values. (a) FeO, (b) CaO, and (c) Al2O3.

Figure 2(a) shows that the FeO activities calculated by the new model are in good agreement with the measured values from Zhou et al. [27]. Meanwhile, our calculation results of the FeO activities are slightly different from Chen’s measurement results [26]. The slight difference might be generated by the differences in thermodynamic values used in the calculation process. The results described in Figure 2(b) and (c) show that the CaO and Al2O3 activities calculated by the model are in good agreement with the results by Chen, which further proved that the model can be used to calculate the activities of slag components accurately.

3.3 Thermodynamic model of desulfurization between slag and metal interface

In the process of establishing the governing equation for calculating the mass action concentration of slag components in the CaF2–CaO–Al2O3–SiO2–TiO2–MgO–FeO system, the structure unit S2− was ignored because the balance number of moles of S2− as a structure unit was very small. Therefore, the calculation of mass action concentration of other structural units was not affected by the number of moles of structured unit S2−.

According to the ion and molecule coexistence theory, only free (Ca2+ + O2−) and (Mg2+ + O2−) ion couples have the ability to desulfurize and participate in the desulfurization reaction. The reaction is expressed as follows:

(12)Ca2++O2+[S]=Ca2++S2+[O]
(13)Mg2++O2+[S]=Mg2++S2+[O]

The equations’ balance constants are expressed as follows:

(14)KCaSθ=αCaSαOαCaOαS=NCaSαONCaOαS=2ω(S)CaS/MS/niαONCaOω[S]fS=ω(S)CaS16NCaOω[S]ni×αOfS
(15)KMgSθ=αMgSαOαMgOαS=NMgSαONCaOαS=2ω(S)MgS/MS/niαONMgOω[S]fS=ω(S)MgS16NMgOω[S]ni×αOfS

The sulfur distribution ratio of (Ca2+ + O2−) and (Mg2+ + O2−) ion pairs can be calculated from equations (14) and (15) as follows:

(16)LS,CaO=ω(S)CaSω[S]=16KCaSθNCaOni×fSαO=16KCaSθNCaOniω[O]×fSfO
(17)LS,MgO=ω(S)MgSω[S]=16KMgSθNMgOni×fSαO=16KMgSθNMgOniω[O]×fSfO

In the electroslag remelting (ESR) process, the oxygen content in the molten metal pool is related to the content of FeO or Al2O3 in the slag [13]. According to the following reactions:

(18)(FeO)=[Fe]+[O]
(19)ΔGθ=1,21,01052.35T(Jmol1)(reference13)
(20)2[Al]+3[O]=Al2O3
(21)ΔGθ=12,02,000+386.30TJmol1(reference14)

The deformations are expressed as follows:

(22)ω[O]=αFeOexp(ΔGFeOθ/ΔRT)fO
(23)RTlnKθ=RTlnαAl2O3αAl2αO3=12,02,000+386.30T
(24)ω[O]=αAl2O3exp(ΔGθ/RT)fAl2ω[Al]2fO313

The activity coefficient of the elements in superalloy can be calculated from equation (25). Because Inconel 718 alloy possesses high alloying element contents, the calculated results of the element activity coefficient could be much more accurate if the secondary interaction coefficient between elements was considered. However, the secondary interactions between elements in nickel-base superalloy activity coefficient have been rarely reported, thus this study only considers the primary activity interaction coefficients [28,29], as shown in Table 7.

(25)logfi=eijω[j]+logω[i]
Table 7

The activity interaction coefficients used in the calculation process

eijNiCrAlMoTiOS
Al−0.0290.00960.045−1.40.03
O0.003−0.0431−0.890.0035−0.6−0.2−0.133

Based on the [Fe]–[O] balance, the oxygen content in alloys is described as:

(26)LS=LS,CaO+LS,MgO=16×KCaSθNCaO+KMgSθNMgO×ni1fO×αFeOexpΔGFeOθ/RT×fsfO

Based on the [Al]–[O] balance, the oxygen content in alloys is described as:

(27)LS=LS,CaO+LS,MgO=16×(KCaSθNCaO+KMgSθNMgO)×niaAl2O3expΔGAl2O3θ/ΔRTfAl2ω[Al]2fO313×fsfO

Based on [Fe]–[Al]–[O] balance, the thermodynamic model of oxygen content in alloys is described as:

(28)LS=LS,CaO+LS,MgO=16×KCaSθNCaO+KMgSθNMgO×ni1fO×αFeOexpΔGFeOθ/ΔRT+αAl2O3expΔGAl2O3θ/ΔRTfAl2ω[Al]2fO313×fsfO

The total sulfur distribution ratio in the slag system can be calculated using equations (26)–(28). The variable fS represents the activity coefficient of dissolved sulfur in the nickel-based alloy. A 1% mass fraction was selected as the standard state, and the sulfur content of our experiment was less than 10 ppm. According to the previous studies of nickel-based alloy solutions, the activity coefficient fS of dissolved sulfur is assumed to be 1. Since the entire experiment process was under the protection of argon atmosphere, and the added superalloy had no oxide layer, the oxygen in the molten metal can only be supplied from the slag. The components in slag that can affect the oxygen content in the molten metal are mainly FeO and Al2O3, and the effects of these two components are independent of each other. So the equilibrium oxygen content based on [Fe]–[Al]–[O] can be calculated as the sum of that of [Fe]–[O] and [Al]–[O] balance. In addition, KCaSθ and KMgSθ can be calculated using equation (29).

(29)Kiθ=exp(ΔGiθ/RT)

where R is the gas constant (8.314 J mol−1 K−1) and T is the absolute temperature (K).

4 Experimental results and discussion

4.1 Experimental results

The sulfur distribution ratio of experimental slags is shown in Table 8. It has been experimentally determined that as the content of FeO and Al2O3 increases, the desulfurization distribution ratio gradually decreases. By changing the content of FeO and Al2O3 in the experiment, slag desulfurization is affected. The effect of FeO will be more significant than that of Al2O3, which mainly affects the desulfurization effect by affecting the slag oxygen content.

Table 8

The sulfur distribution ratio of experimental slags

Slag systemS0S1S2S3S4S5S6S7S8
ln Ls4.93164.17073.91483.77074.00683.85953.72213.60443.5441

4.2 Verification of the desulfurization model

Figure 3 shows the effect of FeO and Al2O3 on the oxygen content in the alloy through theoretical calculation based on ion and molecule coexistence theory and experimental data. It is obvious that the increase in FeO and Al2O3 will increase the oxygen content in the alloy, and the influence of FeO is obviously much greater than that of Al2O3.

Figure 3 Influence of FeO and Al2O3 on the oxygen content in the alloy.
Figure 3

Influence of FeO and Al2O3 on the oxygen content in the alloy.

Taking the S4 slag system as an example, the calculated oxygen contents according to [Fe]–[Al]–[O] balance, [Fe]–[O] balance, and [Al]–[O] balance are shown in Figure 4. It can be seen from the figure that the influence of [Fe]–[O] balance on oxygen content is much greater than that of [Al]–[O] balance, but the influence of [Al]–[O] balance cannot be ignored. The [Fe]–[Al]–[O] balance is more comprehensive in characterizing the oxygen content. Figure 5 shows the comparison of sulfur distribution ratios at different FeO contents obtained through three characterization methods and experiments. In Figure 5, Ls,[Fe]–[O],calculated line is the sulfur distribution ratio calculated based on the desulfurization thermodynamic model established by [Fe]–[O] balance; Ls,[Al]–[O],calculated line is the calculated sulfur distribution ratio based on [Al]–[O] balance; Ls,[Fe]–[Al]–[O],calculated line is the calculated sulfur distribution ratio based on [Fe]–[Al]–[O] balance; Ls,tested line is the sulfur distribution ratio of slag system S0, S1, S2, and S3 measured by experiment. The sulfur distribution ratio determined by any characterization method or experiment decreases with the increase in FeO content at the basicity of 6.67. When the content of FeO in the slag is less than 0.4%, the variation ratio of sulfur distribution is larger. When the FeO content is between 0.4% and 0.7%, the variation range of sulfur distribution ratio is significantly reduced. With the increase in FeO content, the effect of FeO content on the sulfur distribution ratio becomes smaller, which is mainly because some FeO in the slag combines with Al2O3, SiO2, and CaO in the slag to form FeO·Al2O3, 2FeO·SiO2, CaO·FeO·SiO2, and so on. The molecule [17] reduces the activity of FeO, so that the sulfur distribution ratio tends to be stable with the change in FeO. In Figure 5, the fluctuation range of Al2O3 content is in the range of 19.96–20.04%, so the change trend of sulfur distribution ratio calculated according to [Al]–[O] balance is not obvious. Line Ls,[Fe]–[O],calculated and Line Ls,[Al]–[O],calculated have an intersection point when the FeO content is 0.014%. On the left side of the intersection point, the sulfur distribution rate calculated based on the [Al]–[O] balance is closer to the experimental value, while the sulfur distribution rate calculated based on [Fe]–[O] balance is closer to the experimental value on the right side of the intersection point. Therefore, we can conclude that the free oxygen released by Al2O3 is higher than the free oxygen released by FeO when the FeO content is less than 0.014%. At this time, the [Al]–[O] balance is better than the [Fe]–[O] balance, but [Fe]–[Al]–[O] balance calculation result is closest to the measured value.

Figure 4 Calculated oxygen contents based on three oxidizability characterization methods.
Figure 4

Calculated oxygen contents based on three oxidizability characterization methods.

Figure 5 Comparison of sulfur distribution ratios at different FeO contents obtained through three characterization methods and experiments.
Figure 5

Comparison of sulfur distribution ratios at different FeO contents obtained through three characterization methods and experiments.

Figure 6 shows comparison of the sulfur distribution ratios at different Al2O3 contents obtained through three characterization methods and experiments. In Figure 6, the Ls,[Al]–[O],calculated line is the sulfur distribution ratio calculated based on the desulfurization thermodynamic model established by [Al]–[O] balance; the Ls,[Fe]–[O],calculated line is the sulfur distribution ratio based on [Fe]–[O] balance; the Ls,[Fe]–[Al]–[O],calculated line is the sulfur distribution ratio based on [Fe]–[Al]–[O] balance; and the Ls,tested line is the sulfur distribution ratio in slag system S1, S4, S5, S6, S7, and S8 measured by the experiment. The slope of line Ls,[Al]–[O],calculated is obviously greater than the line Ls,[Fe]–[O] and line Ls,[Fe]–[Al]–[O], which indicates that the effect of [Al]–[O] balance increases gradually with the increase in Al2O3. This further illustrates that when the Al2O3 content is higher than 20%, the decomposition of Al2O3 has a greater impact on desulfurization, which cannot be ignored. It can still be seen from the figure that the sulfur distribution ratio obtained by [Fe]–[Al]–[O] balance is closest to the experimental value.

Figure 6 Comparison of sulfur distribution ratios at different Al2O3 contents obtained through three characterization methods and experiments.
Figure 6

Comparison of sulfur distribution ratios at different Al2O3 contents obtained through three characterization methods and experiments.

From the analysis of the reaction equation of desulfurization, it can be known that when the slag shows low oxidizability, it will promote the desulfurization reaction. The FeO content in the slag will directly affect the dissolved oxygen content, which in turn affects desulfurization. The experimental and theoretical calculation values of the sulfur distribution ratio in the slag decrease with the increase in the FeO content. When the FeO content is higher than 0.014%, Al2O3 will play a weak role. When the FeO content is lower than 0.014%, Al2O3 will play a stronger role than FeO. As the content of Al2O3 increases gradually, the sulfur distribution ratio changes greatly. However, the sulfur distribution ratio calculated by the desulfurization thermodynamic models has a large deviation from the experimental value, and the theoretical calculation values are higher than the experimental measurement value. In addition, during the experiment, the S0 slag which does not contain FeO still has desulfurization ability and a large sulfur distribution ratio. However, the desulfurization thermodynamic model based on the [Fe]–[O] balance cannot explain this phenomenon. This indicates that the dissolved oxygen content in the sample is affected not only by the FeO content but also by other factors such as Al2O3. Taniguchi et al. [13] has used the activity of FeO to characterize the dissolved oxygen content in the steelmaking process, but the calculated values of the model and the experimentally determined values are quite different, which further confirms this view. Based on the desulfurization thermodynamic model established by [Fe]–[Al]–[O] balance, Ls,[Fe]–[Al]–[O],calculated is lower than Ls,[Fe]–[O],calculated and closer to the experimental measurements. In addition, when the FeO content is 0, Ls,[Fe]–[Al]–[O],calculated is 746.03, and this model can explain the phenomenon that the S0 slag system has a large sulfur distribution ratio. And as shown in Figure 6, this also verifies that Ls,[Fe]–[Al]–[O],calculated is closer to the measured value. Therefore, using the desulfurization thermodynamic model established by [Fe]–[Al]–[O] balance is more reasonable and feasible compared to other models. However, the sulfur distribution ratio calculated by [Fe]–[Al]–[O] balance is still higher than that of the experimental value. This may be due to the insufficient protection of the atmosphere during the experiment. The element Fe in the alloy is oxidized, making the actual FeO content higher than the FeO content used in the model calculation, resulting in a large deviation. It can be seen from Figure 6 that when the FeO or Al2O3 content is high, the gap between the model calculated value and the experimental value becomes smaller.

When smelting ultra-pure alloys during electroslag remelting, the sulfur content of the alloy must be strictly controlled. Therefore, the content of FeO and Al2O3 must be strictly controlled to be low. In addition, the slag temperature in the electroslag remelting process is usually higher than 1,923 K, and that Al2O3 content in the electroslag remelted slag system usually exceeds 25%, so it is necessary to consider the effect of the oxygen released by the decomposition reaction of Al2O3 on desulfurization. The desulfurization thermodynamic model established by [Fe]–[Al]–[O] balance is more accurate than the desulfurization thermodynamic model constructed by [Fe]–[O] and [Al]–[O] balance; therefore, this method is more suitable for oxygen content characterization in the process of electroslag remelting. Superalloys have very strict requirements for oxygen, sulfur, and other components. Generally, it is required to control the oxygen content to less than 0.0012% and the sulfur content to less than 0.0020%. Therefore, FeO must be strictly controlled during the electroslag remelting process. According to the calculation results in Figure 5, it is recommended to control the FeO content to less than 0.014%.

5 Conclusion

  1. The desulfurization thermodynamic model established by [Fe]–[Al]–[O] balance to characterize slag oxidizability is more accurate compared to the models established by the [Fe]–[O] balance and [Al]–[O] balance, and it is closest to the experimental results.

  2. When the FeO content in the slag is greater than 0.014%, the effect of FeO on the slag sulfur distribution ratio is significantly higher than Al2O3. When the FeO content in the slag is less than 0.014%, the effect of Al2O3 on slag oxidizability will be greater than that of FeO. It is recommended to control FeO content to less than 0.014% when smelting superalloys.

Symbol list

ai

Activity of components i;

bi

Mole number of components i in 100 g of slags (mol);

ƒi

Activity coefficient of component i in nickel-base alloy;

Kiθ

Chemical balance constant of forming component i or structural unit i;

LS

Sulfur distribution ratio between slags–metal;

ni0

Mole number of components i in 100 g of slags (mol);

ni

balance mole number of structural unit i or ion couple i (mol);

R

Gas constant (8.314 J mol−1 K-1);

T

Absolute temperature (K);

ΔrGm,ciθ

Standard molar Gibbs free energy of complex molecule i (J mol−1);

ni

Total balance mole number of all structural units (mol);

ω(i)

Mass percentage of component i in slags (%);

ω[i]

Mass percentage of component i in nickel-base alloy (%);

ω(FeO)

Mass percentage of FeO in slags (%);

ω(Al2O3)

Mass percentage of Al2O3 in slags (%);

ω(S)CaS

Sulfur content in slags boned as CaS (%);

ω(S)MgS

Sulfur content in superalloy boned as MgS (%);

Ls,[Fe]–[Al]–[O],calculated

Calculated sulfur distribution ratio based on [Fe]–[Al]–[O] balance;

Ls,[Fe]–[O],calculated

Calculated sulfur distribution ratio based on [Fe]–[O] balance;

Ls,tested

Tested total sulfur distribution ratio between slag and metal.

Acknowledgments

The authors gratefully express their appreciation to Natural Science Foundation of China (No. 51974153, U1960203), The Joint Fund of State key Laboratory of Marine Engineering and University of Science and Technology Liaoning (No. SKLMEA-USTLN-201901, SKLMEA-USTL-201707).

References

[1] Takasugi, T., and D. P. Pope. Effects of sulfur and phosphorus on the creep ductility of a Cr MoV steel. Materials Science & Engineering, Vol. 57, No. 1, 1983, pp. 15–20.10.1016/0025-5416(83)90021-6Search in Google Scholar

[2] Hu, Z. Q., W. R. Sun, and S. R. Guo. Effect of P, S and Si on the solidification, segregation, microstructure and mechanical properties in Fe–Ni base superalloys. Acta Metallurgica Sinica (English Letters), Vol. 9, No. 6, 1996, pp. 443–452.Search in Google Scholar

[3] Dong, J. X., X. S. Xie, and R. G. Thompson. The influence of sulfur on stress-rupture fracture in Inconel 718 superalloys. Metallurgical & Materials Transactions A, Vol. 31, No. 9, 2000, pp. 2135–2144.10.1007/s11661-000-0131-1Search in Google Scholar

[4] Utada, S., Y. Joh, M. Osawa, T. Yokokawa, T. Sugiyama, T. Kobayashi, et al. Creep property and phase stability of sulfur-doped Ni-base single–crystal superalloys and effectiveness of CaO desulfurization. Metallurgical & Materials Transactions A, Vol. 49, 2018, p. 4029.10.1007/s11661-018-4710-4Search in Google Scholar

[5] Bricknell, R. H., R. A. Mulford, and D. A. Woodford. The role of sulfur in the air embrittlement of nickel and its alloys. Metallurgical Transactions A, Vol. 13, No. 7, 1982, pp. 1223–1232.10.1007/BF02645505Search in Google Scholar

[6] Wu, H. Y., Z. M. Fu, Q. Q. Zhang, M. Y. Li, R. Wei, Y. Cao, et al. Effect of sulfur on the hot plasticity of GH625 superalloy. Materials Science Forum, Vol. 849, 2016, pp. 537–541.10.4028/www.scientific.net/MSF.849.537Search in Google Scholar

[7] Sun, W. R., S. R. Guo, D. Z. Lu, and Z. O. Hu. Effect of sulfur on the solidification and segregation in Inconel 718 alloy. Materials Letters, Vol. 31, No. 3, 1997, pp. 195–200.10.1016/S0167-577X(96)00266-2Search in Google Scholar

[8] Guo, H., M. C. Chaturvedi, and N. L. Richards. Effect of sulphur on hot ductility and heat affected zone microfissuring in Inconel 718 welds. Science & Technology of Welding & Joining, Vol. 5, No. 6, 2000, pp. 378–384.10.1179/136217100101538443Search in Google Scholar

[9] Li, X., X. Geng, and Z. H. Jiang. Influences of slag system on metallurgical quality for high temperature alloy by electroslag remelting. Iron and Steel, Vol. 50, No. 9, 2015, pp. 41–46 (in Chinese).Search in Google Scholar

[10] Geng, X., Z. H. Jiang, F. B. Liu, and L. K. Lian. Control of oxygen content in ESR process. Journal of Materials and Metallurgy, Vol. 8, No. 1, 2009, pp. 16–20 (in Chinese).Search in Google Scholar

[11] Hou, D., Z. H. Jiang, Y. W. Dong, Y. Li, W. Gong, and F. B. Liu. Mass transfer model of desulfurization in the electroslag remelting process. Metallurgical & Materials Transactions B, Vol. 48, No. 3, 2017, pp. 1–13.10.1007/s11663-017-0921-0Search in Google Scholar

[12] Liu, S., H. Lu, H. J. Guo, S. C. Duan, and W. S. Yang. Activity of FeO in the hexabasic slags of electroslag remelting. Journal of Materials and Metallurgy, Vol. 16, No. 1, 2017, pp. 30–37 (in Chinese).Search in Google Scholar

[13] Taniguchi, Y., K. Morita, and N. Sano. Activities of FeO in CaO–Al2O3–SiO2–FeO and CaO–Al2O3–CaF2–FeO slags. Transactions of the Iron & Steel Institute of Japan, Vol. 37, No. 10, 1997, pp. 956–961.10.2355/isijinternational.37.956Search in Google Scholar

[14] Chang, L. Z., X. F. Shi, and J. Q. Cong. Study on mechanism of oxygen increase and countermeasure to control oxygen content during electroslag remelting process. Ironmaking & Steelmaking, Vol. 41, No. 3, 2014, pp. 182–186.10.1179/1743281213Y.0000000114Search in Google Scholar

[15] Yang, X. M., M. Zhang, C. B. Shi, G. M. Chai, and J. Zhang. A sulfide capacity prediction model of CaO–SiO2–MgO–FeO–MnO–Al2O3 slags during the LF refining process based on the ion and molecule coexistence theory. Metallurgical & Materials Transactions B, Vol. 43, No. 2, 2012, pp. 241–266.10.1007/s11663-011-9612-4Search in Google Scholar

[16] Duan, S. C., X. Shi, M. C. Zhang, B. Li, W. S. Yang, F. Wang, et al. Effect of slag composition on the deoxidation and desulfurization of Inconel 718 superalloy by ESR type slag without deoxidizer addition. Metallurgical and Materials Transactions B, Vol. 51, No. 1, 2020, pp. 353–364.10.1007/s11663-019-01729-3Search in Google Scholar

[17] Duan, S. C., X. L. Guo, and H. J. Guo. A manganese distribution prediction model for CaO–SiO2–FeO–MgO–MnO–Al2O3 slags based on IMCT. Ironmaking & Steelmaking: Processes, Products and Applications, Vol. 44, No. 3, 2016, pp. 1–17.10.1080/03019233.2016.1198859Search in Google Scholar

[18] Hou, D., Z. Jiang, Y. Dong, Y. Cao, H. Cao, and W. Gong. Thermodynamic design of electroslagremelting slag for high titanium and low aluminium stainless steel based on IMCT. Ironmaking & Steelmaking, Vol. 43, No. 7, 2016, pp. 1–9.10.1080/03019233.2015.1110920Search in Google Scholar

[19] Shi, C. B., X. M. Yang, J. S. Jiao, C. Li, and H. J. Guo. A sulphide capacity prediction model of CaO–SiO2–MgO–Al2O3 ironmaking slags based on the ion and molecule coexistence theory. ISIJ International, Vol. 50, No. 10, 2010, pp. 1362–1372.10.2355/isijinternational.50.1362Search in Google Scholar

[20] Yang, X. M., M. Zhang, J. L. Zhang, P. C. Li, J. Y. Li, and J. Zhang. Representation of oxidation ability for metallurgical slags based on the ion and molecule coexistence theory. Steel Research International, Vol. 85, No. 3, 2014, pp. 347–375.10.1002/srin.201200138Search in Google Scholar

[21] Li, S. J., G. G. Cheng, L. Yang, L. Chen, Q. Z. Yan, and C. W. Li. A thermodynamic model to design the equilibrium slag compositions during electroslag remelting process: description and verification. ISIJ International, Vol. 57, No. 4, 2017, pp. 713–722.10.2355/isijinternational.ISIJINT-2016-655Search in Google Scholar

[22] Chen, J. X. Commonly used chart data manual for steelmaking. Metallurgical Industry Press, Beijing, China, 1984.Search in Google Scholar

[23] Yang, X. M., J. S. Jiao, R. C. Ding, C. B. Shi, and H. J. Guo. A thermodynamic model for calculating sulphur distribution ratio between CaO–SiO2–MgO–Al2O3 ironmaking slags and carbon saturated hot metal based on the ion and molecule coexistence theory. ISIJ International, Vol. 49, No. 12, 2009, pp. 1828–1837.10.2355/isijinternational.49.1828Search in Google Scholar

[24] Zhang, J. Coexistence Theory of slag structure and its application to calculation of oxidizing capability of slag melts. Iron and Steel Research International, Vol. 10, No. 1, 2003, pp. 1–10.Search in Google Scholar

[25] Zhang, J. Computational thermodynamics of metallurgical melts and solutions. Metallurgical Industry Press, Beijing, China, 2007.Search in Google Scholar

[26] Chen, L., Q. Z. Yan, C. W. Li, M. Liang, S. J. Li, M. L. Qu, et al. Influence of basicity of slag in electroslag remelting furnace on variation of Al-Si content in G20CrNi2Mo Bearing steel, Proceedings of the 2016 National Symposium on Special Metallurgy, Dalian, China, 2016, pp. 305-309 (in Chinese).Search in Google Scholar

[27] Zhou, D. G., W. G. Xu, P. Wang, J. Fu, J. H. Xu, C. S. Wang, et al. Control and behavior of oxygen during electroslag remelting of bearing steel. Iron and Steel, Vol. 33. No. 3, 1998, pp. 13–17 (in Chinese).Search in Google Scholar

[28] Chen, X. C., F. Wang, C. B. Shi, H. Ren, and D. Feng. Effect of electroslag remelting process on desulphurization of GH4169. Journal of Iron and Steel Research, Vol. 24, No. 12, 2012, pp. 11–16 (in Chinese).Search in Google Scholar

[29] Kong, L. Z., Z. Y. Deng, L. Cheng, and M. Y. Zhu. Reaction behaviors of Al-Killed medium manganese steel with glazed MgO refractory. Metallurgical and Materials Transactions B, Vol. 49, No. 6, 2018, pp. 3522–3533.10.1007/s11663-018-1390-9Search in Google Scholar

Received: 2020-03-24
Revised: 2020-06-20
Accepted: 2020-07-29
Published Online: 2020-09-10

© 2020 Wanming Li et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.4.2024 from https://www.degruyter.com/document/doi/10.1515/htmp-2020-0084/html
Scroll to top button