Next Article in Journal
Vibroacoustic Monitoring Features of Radiation-Beam Technologies by the Case Study of Laser, Electrical Discharge, and Electron-Beam Machining
Previous Article in Journal
On the Unique Microstructure and Properties of Ultra-High Carbon Bearing Steel Alloyed with Different Aluminum Contents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Behavior of Nitrogen in GH4169 Superalloy Melt during Vacuum Induction Melting Using Returned Materials

1
State Key Laboratory of Advanced Metallurgy, University of Science and Technology Beijing, Beijing 100083, China
2
AVIC Shangda, Xingtai 054800, China
*
Author to whom correspondence should be addressed.
Metals 2021, 11(7), 1119; https://doi.org/10.3390/met11071119
Submission received: 31 May 2021 / Revised: 5 July 2021 / Accepted: 6 July 2021 / Published: 13 July 2021

Abstract

:
The nitrogen behavior of superalloy melt GH4169 during the vacuum induction melting (VIM) process was clarified by using different proportions of returned materials including block-shaped returned material, chip-shaped returned material, and pure materials to produce a high–purity superalloy melt and provide guidance for the purification of the superalloy melt. For the nitrogen removal during the VIM process, the denitrification rate in the refining period reached 10 ppm per hour on average, which is significantly higher than 1 ppm per hour on average in the melting period. The denitrification reaction of superalloy melt GH4169 under extremely low vacuum pressure is controlled by both the mass transfer of nitrogen in the melt and the chemical reaction of the liquid–gas interface. The nitrogen removal of superalloy melts during VIM occurs through the two methods of gasification denitrification and nitride floatation because the nitrides begin to precipitate in the liquid phase at 1550 °C. A higher nitrogen removal rate can be obtained by increasing the proportion of chip-shaped material or decreasing the proportion of block-shaped material.

1. Introduction

The Ni-based superalloy has been widely applied in the hot parts of aerospace turbine engines and chemical equipment [1]. The trace elements in superalloys can be classified into beneficial elements (e.g., C, B, Zr, Mg, Re, etc.) and harmful elements (e.g., H, O, N, S, Se, Te, Sb, Bi, Pb, Cu, etc.) [2]. Most of the harmful elements such as H, Se, Te, Bi, Pb, and Cu can be easily removed in the form of gas during the vacuum induction melting (VIM) process because of their low solubility in the melt or high vapor pressure. However, the highly harmful oxygen and nitrogen are difficult to remove because these impurities can easily contaminate the melt via the charge, refractory materials, and environmental gases. The removal and purification of oxygen, nitrogen and sulfur in the superalloy, and the control of segregation and precipitate, are gradually becoming the hotspots of superalloys [3,4,5,6,7].
The mechanical properties of a superalloy are influenced by the characteristics, size and morphologies of inclusions and carbides. Inclusions degrade the mechanical properties (e.g., tensile strength and fracture, specifically fatigue properties) of superalloys [8,9,10,11,12]. It was reported in [13] that the single mixed inclusions (oxide as the core, nitride as the shell) in the slab were mainly transformed to clusters during levitation melting of an Ni-based supper alloy. The turbulent collision dominates the aggregation of inclusions, while Stokes collision dominates the floatation of inclusions. The morphologies of carbonitrides in Inconel 718 superalloy were changed from cluster block or single octahedral in ingots to skeleton-like after calcium treatment in the electroslag remelting (ESR) process due to the modification of oxide inclusions by Ca-treatment resulting in a change in precipitation and growth conditions for carbonitrides [14]. Song et al. [15] found that La added in a 22Cr-14W-2Mo superalloy increases the heterogeneous nucleus of M6C carbides and refines the carbides in the superalloy. It was reported in [16,17] that rare-earth elements can promote the phase transformation, counteract deleterious elements such as N, O, S, P, etc., and refine the microstructure of the superalloy due to their special physical and chemical characteristics [18]. Nitrogen and sulfur present in superalloys are treated as impurities in the charge materials and in each step of the alloy processing because they have a deleterious effect on the adherence of the oxide layer, which is a critical feature for the life span of the superalloy [19]. Ta-addition to Ti-containing Ni-based alloys can promote external alumina formation by forming a mixed Ti/Ta oxide compound, hence preventing the enhancement of chromia growth by Ti incorporation, which makes a positive contribution on oxidation resistance in SO2 containing, high pO2 environments [20].
In the practical production of nickel-based superalloys, castings with serious defects, such as sprues, runners, risers, etc., will inevitably be produced. These castings cannot be directly employed for the use in parts manufacturing. In addition, waste materials such as chips, corners and scrap parts are also generated during the processing [21,22]. The final weight of finished parts often only accounts for 30% of the weight of the original smelted alloy and even only 10% for some parts with complex shapes. More than 70% of the production of superalloys is in the form of sprues, runner, risers, scrap parts and cutting pieces. These scraps are similar in composition to the final product of the superalloy and can be used as returned materials to produce superalloy on the premise that harmful elements included in the returned materials can be well controlled.
For superalloys, nitrogen exists in solution or as nitride or carbonitride. Coarse primary TiN inclusions occur when the nitrogen content is higher than the saturated solubility of TiN at solidus. The content of such inclusions in the alloy is more than one order of magnitude larger than that of oxide inclusions, which will seriously affect the mechanical behavior of nickel-based superalloys [23]. In addition, the presence of Ti, Nb, and other elements with high affinity to nitrogen further increases the difficulty of nitrogen removal. With the continuous development of material science, superalloys require a strict control range of nitrogen content, which must be further reduced [24]. Therefore, it is important to study the denitrification reaction mechanism and nitride formation rule of the superalloy during vacuum smelting for further realization of low nitrogen control. The present work aimed to clarify the nitrogen behavior in the superalloy melt when returned materials are used to produce superalloy melt, and to provide guidance for the purification of superalloy melt.

2. Experimental Process and Methods

2.1. Material Preparation

Considering the influence of impurities and harmful elements in the returned materials on the cleanness of the superalloy, the returned materials must be cleaned before charging into the furnace. For the block-shaped returned material, the oxide layer of the surface was peeled off by machining, while the chip-shaped returned material was chopped and cleaned with a special cleaning agent (the components of which are confidential) as shown in Figure 1.
Three cases, listed in Table 1, were designed to investigate the behavior of nitrogen in superalloy melt during vacuum induction melting using different ratios of block-shaped and chip-shaped returned materials and pure materials. Block-shaped returned material (BRM) is mainly from the recycling of sprues, runners, risers, and scrapped superalloy machine parts. Chip-shaped returned material (CRM) is mainly from chips in the cutting and machining of the forging. The composition and cleanness of CRM are close to the final product, but the fluid oil adhering on the surface of CRM will cause the pickup of harmful elements during the melting process. Pure materials (PM) are high-purity main element alloys with very low residual elements.
In the three cases, the pure material was maintained at the same level of 30 wt%, and the returned materials in different cases were changed. Block-shaped and chip-shaped returned materials were used in the Case-A and Case-B, respectively. In Case-C, the mass ratio of block-shaped and chip-shaped returned material was 5:9.

2.2. Experimental Procedure

The industrial production route of GH4169 superalloy is ”vacuum induction melting (VIM) → electroslag remelting (ESR) → vacuum arc remelting (VAR)”. Active metal elements (such as Ti) in GH4169 can easily be oxidized and nitrided when they are melted in the atmosphere, forming a large amount of inclusion. High quality products can only be obtained by vacuum smelting. VIM can greatly remove the harmful elements and gas in the melt through the excellent vacuum atmosphere, guaranteeing the precise chemical composition in the ingot, but it is difficult to eliminate the metallurgical defects (e.g., loose, shrinkage cavity, segregation) in the ingot. ESR is very efficient for deep desulfurization and inclusion removal, but macro-segregation is easily formed in the center of the ingot of ESR. VAR possesses the merit of refining the grain and improving the segregation, which guarantees the hot-machining performance of the superalloy. The process combination of VIM, ESR, and VAR provides the conditions for smelting a high-performance superalloy with returned materials.
The present study focused on the degassing and purification of melt during the VIM process. The superalloy was smelted in a 3-ton vacuum induction melting furnace (VIM-VIDP1000 manufactured by ALD in Hanau, Germany) with a certain ratio of returned materials (70%) and pure material (30%), and the total charging capacity of the alloy was about 2.3 ton per furnace. The air leak rate of the furnace was 32 Pa·L·s−1. The ultimate vacuum degree of the VIM can reach 0.07 Pa, which is very beneficial for degassing and removing inclusions. The target chemical composition of superalloy GH4169 is listed in Table 2 (T.O and T.N represent the sum of free nitrogen and oxygen in metals and combined nitrogen and oxygen). In the smelting process, the target composition was based on the median value of the standard chemical composition with the consideration of effectively reducing the contents of harmful elements of oxygen and nitrogen in the melt to below 10 ppm and 55 ppm, respectively.
During the VIM process, the operation timing of the charge materials addition directly affects the loss of elements and the purification effect of the melt. Therefore, elements that are not easily oxidized and volatilized were charged before the induction heating, while the oxidized and volatile elements were added successively after the raw material was melted. Figure 2 shows the detailed operations during the VIM process. In order to reduce the loss caused by oxidation during the heating process, the vacuum pressure in the smelting chamber must be decreased to below 30 Pa before the raw material was charged. Then, the charge materials including pure nickel-plate (Ni > 99.987 wt%), pure iron (Fe > 99.98 wt%), molybdenum bar (Mo > 99.99 wt%), and returned materials (e.g., block-shaped material and chip-shaped material) were gradually added into the crucible in batches according to the target composition. The induction furnace started the heating operation to melt the charged materials after the first batch of about 200 kg of charge materials were added. The total melting time was about 5 h because the charge material was added eight times at 15 Pa pressure. The melting period lasted for about 5 h under the pressure of 15 Pa. At the end of the melting period, the nickel–niobium alloy (Ni + Nb > 99.3%) and pure chromium (>99.4%) were added into the melt for alloying. When the charging material was fully melted, Sample-1 was taken after 5 min using a vacuum sampler.
After the vacuum pressure was decreased from 15 Pa to below 1 Pa, the process turned to the refining period. The alloying of aluminum and titanium and the fine adjustment of chemical composition were carried out based on the chemical composition of Sample-1, and the samples were taken at intervals of 30 min until the end of the refining period. When the target chemical composition of the superalloy melt was reached, the sample-5 was taken before the casting. Subsequently, the superalloy melt was poured into an ingot mould with Pa of protective argon atmosphere; meanwhile filling argon into the smelting chamber was applied to decrease the reoxidation of the casting process. The process of argon flushing and charging lasted about 30 min in the pouring period, and the actual pure pouring time was about 20–30 min.
All the samples were taken through the automatic sampling device fixed on the top of the smelting chamber. Specimens, including two rod-shaped specimens with ϕ 5 mm ×15 mm, one cubic-shaped specimen with 10 mm × 10 mm × 10 mm, and one cylinder-shaped specimen with size of ϕ 30 mm × 30 mm (height), were machined from the middle part of the samples for the analysis of nitrogen, inclusions, and chemical compositions after a standard metallographic process.
The nitrogen of the samples was analyzed using an automatic analyzer of oxygen and nitrogen (TCH 600, LECO, St. Joseph, MI, USA). The chemical compositions of the samples were analyzed via a Direct Reading Spectrometer (ARL iSpark 8860, Thermo Fisher, Waltham, MA, USA). The sizes, shapes, and chemical composition of the inclusions were characterized by an automatic inclusions analysis system (Evo18, Zeiss, Oberkochen, Germany), with a scanning area of 25 mm2. The formation of inclusions and degassing of the superalloy melt under vacuum atmosphere were analyzed by using thermodynamic software and relevant databases (FactSage 7.2, codeveloped by Thermfact/CRCT, Montreal, QC, Canada, and GTT-Technologies, Herzogenrath, Germany).

3. Results and Discussion

3.1. Operation Curves and Chemical Composition of Superalloy Melt during Smelting Process

The smelting process of the superalloy includes three stages: melting period (300 min), refining period (150 min), and casting period (60 min). Figure 3 shows the operation curves of temperature and pressure at different stages. During the melting period, the materials were charged into the furnace in batches under a pressure of below 15 Pa until all the materials were fully melted, and the temperature changed between 1775 and 1800 K. When the pressure decreased from 12 Pa to below 1 Pa, the refining period started. The temperatures in the refining period were maintained between 1775 and 1800 K. After 150 min refinement, the high-purity superalloy melt was poured into an ingot mold with Pa of protective argon atmosphere, and the temperature of the casting process was controlled between 1755 and 1745 K. Table 3 shows the chemical composition of different cases during the smelting process.

3.2. Change of Nitrogen in the Superalloy Melt during Smelting Process

As shown in Figure 4, the square-, circle-, and triangle-marked curves represent Cases-A, -B, and -C, respectively. It can be seen that in Case-C two data points are missing due to the sampling failure. The initial nitrogen of the charge materials changed between 70 and 80 ppm regarding different combinations of returned materials estimated by the calculation of materials charged and the initial chemical compositions in the raw materials. Then, it decreased to the range of 61 to 64 ppm after all the charge materials had been melted. The difference in nitrogen content for different returned materials narrowed at the melting endpoint. The denitrification rate of the melt maintained a very low level during the 5 h of the melting process, because the charge materials were added in batches to melt them gradually to avoid splashing. The denitrification rate in the refining period reached 10 ppm per hour on average, which is significantly higher than 1 ppm per hour on average in the melting period. Low vacuum pressure (<1 Pa) and good stirring conditions promote the removal of nitrogen in the melt. Despite the protection of argon atmosphere, the melt still had a significant nitrogen pickup due to air leakage during the casting process. Case-B obtained the best denitrification effect and lowest nitrogen in the final product.
Figure 5 was obtained by the statistical treatment of the number density and average size of inclusions in the EVO results, and shows the change of quantity density and average size of nitride in the superalloy samples obtained during the refining process. In Figure 5, there is a difference in the number of nitrides in Case–A, –B, and –C. This difference may be caused by the difference in the number of oxides due to the bimodal effect of charging during the smelting period [25]. The surface area of different shapes of charge materials is different, which may lead to different oxide films brought in by them and may lead to different nitrogen compounds in the initial stage of refining [26].
With the refining time, the quantity density of nitrides in different cases shows a similar trend, changing from the average 2.1 per mm2 at the melting endpoint to 1.1 per mm2 at the refining end to 1.3 per mm2 at the casting end. The trend of the variation in the quantity density of nitrides is consistent with that of the nitrogen content, i.e., as the nitrogen content continues to decrease, the quantity density of nitrides also decreases correspondingly. By comparison of the three cases, Case-C with 25% BRM plus 45% CRM plus 30%PM has the largest quantity density of nitride in the casting end, reaching 1.4 per mm2. The quantity density of nitrides decreases with the refining time, but the average size changes little, and remains in a range of 1 to 1.5 μm. The single returned material cases (Case-A and Case-B) can obtain lower total nitrogen than the mixed returned material case (Case-C).

3.3. Correlation of Denitrification and Nitrogen Solubility in the Superalloy Melt

Nitrogen in the superalloy exists in three forms: solid solution, compound, and gas, depending on the nitrogen content in the superalloy melt and its saturated solubility. The solubility of nitrogen in the melt and the partial pressure of nitrogen in the environment obey Sievert’s law [27,28]. The Gibbs free energy of reaction, as shown in Equation (1), can be calculated by Equation (2):
1 2 N 2 ( g ) = [ N ]
lgK N = 3630 T 0.883
where, K is the equilibrium constant expressed as K N = f N [ % N ] ( p N / p θ ) 0.5 ; [%N] is the dissolved nitrogen in the melt, wt%; PN is the dimensionless nitrogen partial pressure of nitrogen; and fN is the activity coefficient of nitrogen in the melt, which can be obtained by the Wagner model using the interaction coefficient of elements in Table 4.
The interaction coefficient of elements under an arbitrary temperature (fN,T) can be calculated through the correlation of lg f T lg f 1873 = 1873 T [33]. Based on the aforementioned equations, the change of nitrogen in the melt with respect to the partial pressure of nitrogen at different temperatures can be obtained, as shown in Figure 6.
It is shown in Figure 6 that, at the end of the melting period, the nitrogen content of 63 ppm on average in the melt is much lower than the theoretical solubility of nitrogen in the melt under the working pressure of 15 Pa. It indicates that the denitrification is difficult to take place under a pressure of 15 Pa after the superalloy is melted. As the pressure decreases to 0.5 Pa on average during the refining period, both nitrogen solubility and nitrogen content in the melt significantly decrease, although there is still a gap between the measured nitrogen in the melt and the theoretical solubility. The phenomenon mentioned above can be attributed to two aspects: (1) the precipitation of nitrogen has not yet reached equilibrium; (2) the nitrogen in the melt is not only in the form of solution. That is, the removal of nitrogen during the smelting process is greatly affected by the kinetic conditions and the chemical composition especially the nitrogen-affinity elements such as Ti, Nb, Al etc., which could form the nitrogen-based compounds.
It was reported in [34,35,36] that denitrification is a first-order reaction when the limiting step is the mass transfer of nitrogen from the interior of the liquid steel to the gas–liquid interface, while it is a second-order reaction when the denitrification rate is dominated by the chemical reaction of nitrogen atoms forming nitrogen molecules at the gas–liquid interface. Thus, the denitrification rate can be described via the first-order reaction model in Equation (3) or via the second-order reaction model in Equation (4). During the smelting process, the gas in the smelting chamber was continuously pumped out and the pressure in the chamber was maintained at a very low level which caused the N2 partial pressure in the gas phase to be extremely low. Here, it is considered that the nitrogen content at the gas–liquid interface is in equilibrium with N2 partial pressure approaching 0.5 Pa. Equations (3) and (4) are integrated, respectively, to obtain Equations (5) and (6). (Y1 and Y2 represent the first-order reaction and second-order reaction equations, respectively.)
d C N d t = A V k l ( C N C N e )
d C N d t = A V k 2 ( C N e 2 C N 2 )
Y 1 = ln C N e C N C N e C N 0 = A V k l t
Y 2 = 1 2 C N e [ ln ( C N C N e C N + C N e ) ln ( C N 0 C N e C N 0 + C N e ) ] = k 2 A V t
where, k1 and k2 are denitrification ratio constants of the first-order reaction and second-order reaction, m·s−1, m·s−1·%−1; CN0 is the initial nitrogen content in the melt, wt%; t is the smelting time, s; A is the surface area of the superalloy melt, m2. V is the volume of the superalloy melt, m3. CN is the nitrogen content in the melt at the time of t, wt%; and CNe is the nitrogen content at the gas/liquid interface in equilibrium with N2 partial pressure, wt%.
The relevant data are derived according to Equations (5) and (6), the fitting result of the reaction order is shown in Figure 7, and the fitting parameters are shown in Table 5. According to the first-order reaction model, the mass transfer coefficient kl is between 7.754 × 10−5 and 9.568 × 10−5 m/s. For the second reaction model, the rate coefficient kc of the interface chemical reaction is between 0.0098 and 0.0127 m∙s−1∙%−1. The slope of the curve increases with the chip-shaped returned material, which indicates high CRM is beneficial for nitrogen removal under extremely low vacuum pressure. The average coefficients of determination, Y1 and Y2, are both above 0.96. It can be deduced that the denitrification reaction of the superalloy melt under extremely low vacuum pressure is controlled by both the mass transfer of nitrogen in the melt and the chemical reaction of the liquid-gas interface.

3.4. Thermodynamic Consideration of Nitride Formation in the Superalloy Melt

During the smelting process of VIM, there are two main sources of nitrogen, i.e., the introduction of nitrogen from the charge materials and the absorption of nitrogen from air caused by the large leakage rate of the smelting chamber. The nitrogen can exist in three forms: solid solution, compound, and gas—depending on the nitrogen content in the superalloy melt and its saturated solubility. When the nitrogen content is lower than the saturation solubility, it presents mainly in the form of solid solution whilethe nitrogen may combine with the nitrogen-affinity elements such as Nb, Ti, Al etc., to form the compound or escape as a gas due to the effect of the vacuum atmosphere.
It can be seen from Figure 8, that the liquidus and solidus temperature are 1355 °C and 1214 °C, respectively. A small amount of the oxides (M2O3: mainly Al2O3, Ti2O3, and other oxides) co-exist with the liquid phase, while the nitride (MN: mainly TiN and (Nb, Mo)N) begins to precipitate in the liquid phase at 1550 °C, reaching the maximum at 1250 °C. The thermodynamic reactions are shown in Equations (7)–(11). It is indicated that many nitrides could already be present in the melt during the degassing process because the high content of nitride-forming elements such as Ti, Nb, Al existed in the returned materials. The precipitation of nitrides in the liquid-phase promotes the dissolved nitrogen in the melt transforming to nitrogen compounds and hinders the nitrogen desorption during the vacuum process. As shown in Figure 9, three typical inclusions, namely single-phase oxide inclusion, two-phase inclusion of titanium nitride wrapping oxide, and multiphase inclusion with three-layer structure, are found in the melt during the refining period. The fine oxide core provides favorable conditions for the precipitation of titanium nitride, which serves as the nucleation core of niobium and molybdenum nitride. The tendency of multiphase nucleation is related to the low degree of mismatch between particles.
2   [ Al ] + 3   [ O ] = Al 2 O 3
2   [ Ti ] + 3   [ O ] = Ti 2 O 3
[ Ti ] + [ N ] = TiN
[ Nb ] + [ N ] = NbN
[ Mo ] + [ N ] = MoN
Figure 10 shows a typical multiphase inclusion with a three-layer structure, i.e., TiN grows on 1 μm of MgO-Al2O3 spinel core, and then (Nb, Mo)N grows on the TiN. This reveals that the formation of precipitates follows the sequence: MgO-Al2O3 spinel, TiN, and (Nb, Mo)N, which can be verified by the phase precipitation sequence in Figure 8.
As shown in Figure 11, the denitrification rate coefficient increases with the amount of Nb and Ti content in the melt. Titanium and niobium are the formation elements of nitride in the melt which can promote the formation of nitrides. It was concluded in Figure 8 that the nitride (MN) begins to precipitate in the liquid phase at 1550 °C, and many nitrides such as TiN and (Nb, Mo)N were formed on the oxides core, as can be seen in Figure 10. The denitrification of the superalloy melt in vacuum process is mainly composed of two parts, one is gasification denitrification depending on the pressure of the vacuum chamber, the other is nitride removal depending on the contents of the formation elements of nitrides in the melt. High initial titanium and niobium contents in the superalloy melt means more nitrides are formed during the smelting process, and a part of these nitrides will be removed by floatation. It is concluded in Table 1 that, during the melting process of returned materials, the initial content of Nb and Ti in different cases present a descending order as follows: Case-B, Case-C, Case-A. It can be seen in Figure 12 that the ratio of titanium and nitrogen in the melt increase with refining time and descends in order of Case-B, Case-C, Case-A. That means that the denitrification rate of the superalloy melt decreased with the proportion increase of the block-shaped returned material during the smelting process. Thus, the block-shaped returned material is not conducive to the denitrification of the superalloy melt.

4. Conclusions

(1)
The denitrification rates of superalloy melt GH4169 during the VIM process were 1 ppm per hour at the melting period and 10 ppm per hour in the refining period. Low vacuum pressure (<1 Pa) and good stirring conditions promote the removal of nitrogen in the melt.
(2)
The denitrification reaction of superalloy melt GH4169 under extremely low vacuum pressure is controlled by both the mass transfer of nitrogen in the melt and the chemical reaction of the liquid–gas interface. The average coefficients of determination for the first-order and second-order reaction are both above 0.96.
(3)
The liquidus and solidus temperature of GH4169 are 1355 °C and 1214 °C, respectively. A small amount of the oxide (M2O3) co-exist with the liquid phase, while the nitride (MN) begins to precipitate in the liquid phase at 1550 °C reaching the maximum at 1250 °C.
(4)
Three typical inclusions, namely single-phase oxide inclusion, two-phase inclusion of titanium nitride wrapping oxide, and multiphase inclusion with three-layer structure, are found in the melt during the refining period.
(5)
The denitrification of the superalloy melt during the vacuum process is due to the gasification denitrification and nitrides floatation, and the denitrification rate decreased with the proportion increase of the block-shaped returned material during the smelting process.

Author Contributions

Conceptualization, M.W. and Y.B.; methodology, S.G. and M.W.; software, X.X. and M.L.; validation, S.G., M.W. and M.L.; formal analysis, M.W. and S.G.; investigation, S.G. and X.X.; resources, M.W.; data curation, S.G., M.L. and M.W.; writing—original draft preparation, M.W. and X.X.; writing—review and editing, M.W. and S.G.; visualization, X.X.; supervision, M.W. and Y.B.; project administration, M.W.; funding acquisition, M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities, grant number FRF-BD-19-022A. The authors wish to express gratitude for the financial support for their research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The experimental materials were provided by the AVIC Shangda, and the experimental data were completed by engineers from the AVIC Shangda. The authors wish to express their gratitude for this necessary help.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lvova, E.; Norsworthy, D. Influence of service-induced microstructural changes on the aging kinetics of rejuvenated Ni-based superalloy gas turbine blades. J. Mater. Eng. Perform. 2001, 10, 299–312. [Google Scholar] [CrossRef]
  2. Zhu, Y.X.; Li, C.; Liu, Y.C.; Ma, Z.Q.; Yu, H.Y. Effect of Ti addition on high-temperature oxidation behavior of Co-Ni-based superalloy. J. Iron Steel Res. Int. 2020, 27, 1179–1189. [Google Scholar] [CrossRef]
  3. Liu, Z.Y.; Zhu, Q. Effect of Pulse Frequency on the Columnar-to-Equiaxed Transition and Microstructure Formation in Quasi-Continuous-Wave Laser Powder Deposition of Single-Crystal Superalloy. Metall. Mater. Trans. A 2021, 52, 776–778. [Google Scholar] [CrossRef]
  4. Liburdi, J.; Lowden, P.; Nagy, D.; de Priamus, T.R.; Shaw, S. Practical Experience with the Development of Superalloy Rejuvenation. In Proceedings of the ASME Turbo Expo, Orlando, FL, USA, 8–12 June 2009; Volume 4, pp. 819–827. [Google Scholar] [CrossRef] [Green Version]
  5. Musavi, S.H.; Davoodi, B.; Niknam, S.A. Environmental-friendly turning of A286 superalloy. J. Manuf. Process. 2018, 32, 734–743. [Google Scholar] [CrossRef]
  6. Shao, Y.L.; Xu, J.; Wang, H. Effect of Ti and Al on microstructure and partitioning behavior of alloying elements in Ni-based powder metallurgy superalloys. Int. J. Miner. Metall. Mater. 2019, 26, 500–506. [Google Scholar] [CrossRef]
  7. Gao, X.Y.; Zhang, L.; Qu, X.H. Effect of interaction of refractories with Ni-based superalloy on inclusions during vacuum induction melting. Int. J. Miner. Metall. Mater. 2020, 27, 1551–1559. [Google Scholar] [CrossRef]
  8. Liu, J.; Zhang, Z.; Li, B.; Lang, S. Multiaxial Fatigue Life Prediction of GH4169 Alloy Based on the Critical Plane Method. Metals 2019, 9, 255. [Google Scholar] [CrossRef] [Green Version]
  9. Xu, R.; Zhou, Y.; Li, X.; Yang, S.; Han, K.; Wang, S. The Effect of Milling Cooling Conditions on the Surface Integrity and Fatigue Behavior of the GH4169 Superalloy. Metals 2019, 9, 1179. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, J.L.; Wei, D.S.; Wang, Y.R.; Zhong, B. High-temperature LCF life estimation based on stress gradient effect of notched GH4169 alloy specimens. Fatigue Fract. Eng. Mater. Struct. 2017, 40, 1640–1651. [Google Scholar] [CrossRef]
  11. Semiatin, S.L.; Tiley, J.S.; Zhang, F.; Smith, T.M.; Zhang, R.Y.; Dong, H.B.; Gadaud, P.; Cormier, J. A Fast-Acting Method for Simulating Precipitation during Heat Treatment of Superalloy 718. Metall. Mater. Trans. 2021, 52, 483–499. [Google Scholar] [CrossRef]
  12. Mueller, E.M.; Carney, L.; Ngin, S.T.; Yadon, J.L. Failure analysis of weld-repaired B-1900 turbine blade shrouds. Superalloys 2008, 2008, 469–477. [Google Scholar] [CrossRef]
  13. Gao, X.Y.; Zhang, L.; Luan, Y.F.; Chen, X.W.; Qu, X.H. Investigation of Inclusion Agglomeration and Flotation during Levitation Melting of Ni-Based Superalloy in a Cold Crucible. JOM 2020, 72, 3247–3255. [Google Scholar] [CrossRef]
  14. Chen, X.C.; Shi, C.B.; Guo, H.J.; Wang, F.; Ren, H.; Feng, D. Investigation of Oxide Inclusions and Primary Carbonitrides in Inconel 718 Superalloy Refined through Electroslag Remelting Process. Metall. Mater. Trans. 2012, 43, 1596–1607. [Google Scholar] [CrossRef]
  15. Song, X.; Wang, L.; Liu, Y.; Ma, H.P. Precipitation characteristics and La effects on precipitates of a new 22Cr-14W-2Mo superalloy. Rare Metals 2010, 29, 132–137. [Google Scholar] [CrossRef]
  16. Pelcová, J.; Smola, B.; Stulíková, I. Influence of processing technology on phase transformations in a rare-earth-containing Mg–Zn–Zr alloy. Mater. Sci. Eng. 2007, 462, 334–338. [Google Scholar] [CrossRef]
  17. Wang, R.M.; Song, Y.G.; Han, Y.F. Effect of rare earth on the microstructures and properties of a low expansion superalloy. J. Alloys Compd. 2002, 33, 575–580. [Google Scholar] [CrossRef]
  18. Guo, F.A.; Xiang, C.J.; Yang, C.X.; Cao, X.M.; Mu, S.G.; Tang, Y.Q. Study of rare earth elements on the physical and mechanical properties of a Cu–Fe–P–Cr alloy. Mater. Sci. Eng. 2008, 147, 1–6. [Google Scholar] [CrossRef]
  19. Gheno, T.; Monceau, D.; Oquab, D.; Cadoret, Y. Characterization of Sulfur Distribution in Ni-Based Superalloy and Thermal Barrier Coatings after High Temperature Oxidation: A SIMS Analysis. Oxid. Met. 2010, 73, 95–113. [Google Scholar] [CrossRef] [Green Version]
  20. Jalowicka, A.; Nowak, W.; Naumenko, D.; Singheiser, L.; Quadakkers, W.J. Effect of nickel base superalloy composition on oxidation resistance in SO2 containing, high pO2 environments. Mater. Corros. 2014, 65, 178–187. [Google Scholar] [CrossRef]
  21. Niu, J.; Sun, X.; Jin, T.; Yang, K.; Guan, H.; Hu, Z. Investigation into deoxidation during vacuum induction melting refining of nickel base superalloy using CaO crucible. J. Mater. Sci. Technol. 2003, 19, 435–439. [Google Scholar] [CrossRef]
  22. Alexander, J. Optimizing deoxidation and desuIphurization during vacuum induction melting of alloy 718. J. Mater. Sci. Technol. 1985, 1, 167–170. [Google Scholar] [CrossRef]
  23. Qian, K.; Chen, B.; Shu, L.; Liu, K. Nitrogen Solubility in Liquid Ni-V, Ni-Ta, Ni-Cr-V, and Ni-Cr-Ta Alloys. Metals 2019, 9, 1184. [Google Scholar] [CrossRef] [Green Version]
  24. Wang, W.; Zhang, L.; Yang, Y.Q.; Gao, M.; Ding, L.L.; Ma, Y.C.; Liu, K. Thermodynamics and Kinetics of Denitrification Reaction of High Cr Nickel Base Alloys. Rare Metal Mater. Eng. 2020, 49, 3803–3808. [Google Scholar]
  25. Campbell, J.; Tiryakioğlu, M. Bifilm Defects in Ni-Based Alloy Castings. Metall. Mater. Trans. 2012, 43, 902–914. [Google Scholar] [CrossRef]
  26. Campbell, J. An overview of the effects of bifilms on the structure and properties of cast alloys. Metall. Mater. Trans. 2006, 37, 857–863. [Google Scholar] [CrossRef]
  27. Niu, J.P.; Yang, K.N.; Sun, X.F.; Jin, T.; Guan, H.R.; Hu, Z.Q. Denitrogenation during vacuum induction melting refining Ni base superalloy using CaO crucible. J. Mater. Sci. Technol. 2002, 18, 1041–1044. [Google Scholar] [CrossRef]
  28. Kowanda, C.; Speidel, M.O. Solubility of nitrogen in liquid nickel and binary Ni–Xi alloys (Xi=Cr, Mo, W, Mn, Fe, Co) under elevated pressure. Scripta Mater. 2003, 48, 1073–1078. [Google Scholar] [CrossRef]
  29. Balasubramanian, K.; Kirkaldy, J.S. Experimental Investigation of the Thermodynamics of Fe–Nb–N Austenite and Nonstoichiometric Niobium Nitride (1373–1673 K). Can. Metall. Quart. 1989, 28, 301–315. [Google Scholar] [CrossRef]
  30. Ohta, H.; Suito, H. Thermodynamics of Aluminum and Manganese Deoxidation Equilibria in Fe-Ni and Fe-Cr Alloys. ISIJ Int. 2003, 43, 1301–1308. [Google Scholar] [CrossRef]
  31. Chen, E.-B.; Wang, S.-J.; Dong, Y.-C.; Wu, B.-G.; Zhou, Y. Thermodynamic Properties of Fe-C-N and Fe-C-B-N Melts. J. Iron Steel Res. Int. 2007, 14, 21–24. [Google Scholar] [CrossRef]
  32. Zhou, L.; Sun, X.F.; Yin, F.; Hou, G.C.; Zheng, Q.; Guan, H.R.; Hu, Z. Effect of melt-superheating treatment on nitrogen-absorption behavior in cast nickel-base superalloy. Chin. J. Nonferrous Met. 2003, 13, 550–553. [Google Scholar] [CrossRef]
  33. Byrne, M.; Belton, G.R. Studies of the interfacial kinetics of the reaction of nitrogen with liquid iron by the15N-14N isotope exchange reaction. Metall. Trans. B 1983, 14, 441–449. [Google Scholar] [CrossRef]
  34. Fruehan, R.J.; Martonik, L.J. The rate of absorption of nitrogen into liquid iron containing oxygen and sulfur. Metall. Trans. B 1980, 11, 615–621. [Google Scholar] [CrossRef]
  35. Ban-Ya, S.; Ishii, F.; Iguchi, Y.; Nagasaka, T. Rate of nitrogen desorption from liquid iron-carbon and iron-chromium alloys with argon. Metall. Trans. B 1988, 19, 233–242. [Google Scholar] [CrossRef]
  36. Wang, Z.D.; Cao, K.J.; He, J.L. Induction Furnace Smelting; Chemical Industry Press: Beijing, China, 2007; 347p. [Google Scholar]
Figure 1. Returned materials before cleaning—(a) chip-shaped and (b) block-shaped—and after cleaning—(c) chip-shaped and (d) block-shaped.
Figure 1. Returned materials before cleaning—(a) chip-shaped and (b) block-shaped—and after cleaning—(c) chip-shaped and (d) block-shaped.
Metals 11 01119 g001
Figure 2. Detailed operations during the process of VIM.
Figure 2. Detailed operations during the process of VIM.
Metals 11 01119 g002
Figure 3. Operation curves of temperature and pressure at different stages.
Figure 3. Operation curves of temperature and pressure at different stages.
Metals 11 01119 g003
Figure 4. Change of nitrogen in the melt with smelting time.
Figure 4. Change of nitrogen in the melt with smelting time.
Metals 11 01119 g004
Figure 5. Change of quantity density and average size of nitride in the superalloy samples obtained.
Figure 5. Change of quantity density and average size of nitride in the superalloy samples obtained.
Metals 11 01119 g005
Figure 6. Nitrogen solubility at different partial pressures and temperatures.
Figure 6. Nitrogen solubility at different partial pressures and temperatures.
Metals 11 01119 g006
Figure 7. Data fitting of denitrification ratio: (a) the first order reaction model and (b) the second-order reaction model.
Figure 7. Data fitting of denitrification ratio: (a) the first order reaction model and (b) the second-order reaction model.
Metals 11 01119 g007
Figure 8. (a) Change of phases fraction in GH 4169 superalloy with temperature, and (b) local enlargement of precipitation phase
Figure 8. (a) Change of phases fraction in GH 4169 superalloy with temperature, and (b) local enlargement of precipitation phase
Metals 11 01119 g008
Figure 9. Typical inclusions in the melt during the refining period.
Figure 9. Typical inclusions in the melt during the refining period.
Metals 11 01119 g009
Figure 10. Mapping of a typical multiphase inclusion with a three-layer structure.
Figure 10. Mapping of a typical multiphase inclusion with a three-layer structure.
Metals 11 01119 g010
Figure 11. Mapping of a typical multiphase inclusion with a three-layer structure.
Figure 11. Mapping of a typical multiphase inclusion with a three-layer structure.
Metals 11 01119 g011
Figure 12. Ratio of titanium and nitrogen in the superalloy melt in the refining period.
Figure 12. Ratio of titanium and nitrogen in the superalloy melt in the refining period.
Metals 11 01119 g012
Table 1. Experimental scheme with different mass ratio of return materials.
Table 1. Experimental scheme with different mass ratio of return materials.
SchemeBlock-Shaped
Returned Material
Chip-Shaped
Returned Material
Pure Material
A70%/30%
B/70%30%
C25%45%30%
A (70%BRM–0%CRM–30%PM); B (0%BRM–70%CRM–30%PM); C (25%BRM–45%CRM–30%PM).
Table 2. Chemical composition of superalloy GH4169, wt%.
Table 2. Chemical composition of superalloy GH4169, wt%.
ElementsCMnNiCrMoAlTiNbFeT.OT.N
Standard range0.015–0.036≤0.3552.0–55.017.00–19.002.80–3.150.30–0.650.75–1.155.20–5.5516.0–19.0≤0.0025≤0.010
Target control0.023/53.518.02.950.60.955.4517.0≤0.001≤0.0055
Table 3. Chemical composition of different cases during the smelting process.
Table 3. Chemical composition of different cases during the smelting process.
CaseTime/MinCMnSiSPNiCrMoAlTiCuBNbCoFe
A
(70%BRM-0%CRM-30%PM)
00.0240.030.0520.00080.0154.6919.012.800.320.680.0180.0035.140.1416.16
600.0230.0230.0530.00080.00953.7518.382.950.560.920.0180.0035.380.2916.74
900.0230.0220.0570.00080.01353.8918.362.970.560.930.0170.0035.430.2916.66
1200.0230.020.0570.00080.01153.7918.362.960.560.920.0170.0035.40.2916.72
1500.0270.0190.0550.00070.00954.2318.342.990.610.990.0160.0035.430.316.71
2000.0290.0180.0540.00070.01153.9718.242.970.60.980.0150.0055.390.316.75
B
(0%BRM-70%CRM-30%PM)
00.0210.0320.0520.00080.00955.4418.532.860.360.850.0260.0035.50.2315.12
600.0240.0190.0530.00080.01353.9617.92.960.60.970.0210.0035.610.316.68
900.0230.0190.0550.00080.01554.1617.932.930.610.950.020.0035.290.316.87
1200.0230.0170.0530.00080.01453.9717.782.960.60.980.0190.0035.640.316.64
1500.0230.0160.0530.00080.0135417.912.960.620.960.0180.0035.60.2916.94
2000.0220.0120.0570.001-53.717.222.950.6340.980.013-5.56-16.9
C
(25%BRM-45%CRM-30%PM)
00.030.0320.0640.0010.0154.8318.472.880.360.640.020.0035.340.1716.7
600.0290.0240.0630.0010.01253.5618.042.970.5720.960.020.0035.520.2917.21
900.0280.0180.0690.0020.01353.7317.92.990.570.960.020.0035.5540.2917.17
2000.0270.0160.0680.00150.01253.417.892.970.580.960.0140.00455.570.317.09
Table 4. Interaction coefficient of the typical elements in the superalloy at 1873 K. Adapted from ref. [29,30,31,32].
Table 4. Interaction coefficient of the typical elements in the superalloy at 1873 K. Adapted from ref. [29,30,31,32].
ElementCCrTiNbMoFe
e N j 0.09−0.1−0.21−0.072−0.043−0.021
Table 5. R2 and Reaction rate k
Table 5. R2 and Reaction rate k
Reaction OrderRaw MaterialR2
First-order reaction model70 CRM + 30 PM0.984
70 BRM + 30 PM0.923
25 BRM + 45 CRM + 30 PM0.992
Second-order reaction model70 CRM + 30 PM0.983
70 BRM + 30 PM0.909
25 BRM + 45 CRM + 30 PM0.984
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gao, S.; Wang, M.; Xie, X.; Liu, M.; Bao, Y. Behavior of Nitrogen in GH4169 Superalloy Melt during Vacuum Induction Melting Using Returned Materials. Metals 2021, 11, 1119. https://doi.org/10.3390/met11071119

AMA Style

Gao S, Wang M, Xie X, Liu M, Bao Y. Behavior of Nitrogen in GH4169 Superalloy Melt during Vacuum Induction Melting Using Returned Materials. Metals. 2021; 11(7):1119. https://doi.org/10.3390/met11071119

Chicago/Turabian Style

Gao, Shengyong, Min Wang, Xiaoyu Xie, Meng Liu, and Yanping Bao. 2021. "Behavior of Nitrogen in GH4169 Superalloy Melt during Vacuum Induction Melting Using Returned Materials" Metals 11, no. 7: 1119. https://doi.org/10.3390/met11071119

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