Next Article in Journal
Analytical Solution for a 1D Hexagonal Quasicrystal Strip with Two Collinear Mode-III Cracks Perpendicular to the Strip Boundaries
Next Article in Special Issue
Plasma-Pulsed GMAW Hybrid Welding Process of 6061 Aluminum and Zinc-Coated Steel
Previous Article in Journal
Spectroscopic Properties of Pb2+-Doped BaF2 Crystals
Previous Article in Special Issue
Precipitation of Topologically Closed Packed Phases during the Heat-Treatment of Rhenium Containing Single Crystal Ni-Based Superalloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Transition of Nb3Sn during the Heat Treatment of Precursors after Mechanical Alloying

1
Institute of Electrical Engineering, Chinese Academy of Sciences, Beijing 100190, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(4), 660; https://doi.org/10.3390/cryst13040660
Submission received: 24 February 2023 / Revised: 28 March 2023 / Accepted: 9 April 2023 / Published: 11 April 2023

Abstract

:
The phase transition process of Nb3Sn during heat treatment exerts important influences on Nb3Sn formation and the superconducting characteristics of Nb3Sn superconductors. A simple method for quickly preparing Nb3Sn was studied. First, Nb, Sn, and Cu powders were mechanically alloyed to prepare the precursor. Then, the precursor was heat treated at different times to form Nb3Sn. During the first stage, the morphology and crystal structure of the products were analyzed after different milling times. The results of the transmission electron microscopy showed the poor crystallinity of the products compared with the original materials. During the second stage, heat treatment was performed at different temperatures ranging from room temperature to 1073 K. After treatment, the products were studied via X-ray diffraction analysis to determine how the structure changed with increasing temperature. Only the Nb diffraction peaks in the precursor were observed after high-energy ball milling for more than 3 h. When the heat treatment temperature was above 773 K and heat treatment time was 15 min, Nb3Sn began to form. When the temperature was above 973 K, some impurities, such as Nb2O5, appeared. After 5 h of ball milling, the precursor was heat treated at different times in a vacuum heat treatment furnace. The crystal structure of the product exhibited evident diffraction peaks of Nb3Sn. The critical temperatures of the samples that were heat treated at different times were between 17 K and 18 K. The magnetic critical current density of the sample versus the applied magnetic field at 4.2 K indicated that the magnetic Jc was approximately 30,000 A/cm2.

1. Introduction

Since Nb3Sn was discovered by Matthias in 1954 [1], research on this compound has elicited considerable attention, particularly regarding its preparation and superconductive characteristics. Compared with NbTi, Nb3Sn has higher critical temperature (above 18 K) and critical magnetic field of around 30 T. Nb3Sn superconductors have important applications in high magnetic field situations, particularly at above 10 T. Current applications of Nb3Sn include magnetic resonance imaging [2], nuclear magnetic resonance [3], dipole and quadrupole magnets for particle accelerators, and other areas where high magnetic fields are required, such as the International Thermonuclear Experimental Reactor [4]. Conventional manufacturing methods, such as bronze technology [5,6], internal tin technology [7], and powder-in-tube [8,9] process, require a long period of heat treatment at around 650 °C to allow Sn to react with Nb and form Nb3Sn superconductors. These methods are time consuming, and their costs are high.
Heat treatment is a key point in the preparation of Nb3Sn. Heat treatment temperature exerts an important effect on the upper critical field Bc2, with a higher reaction temperature leading to higher Bc2 [10]. Many studies on heat treatment have investigated the influences of temperature variation, heat treatment time, temperature gradient, ascending temperature speed, and temperature uniformity on the superconducting characteristics of Nb3Sn [11]. Various factors that are required to form fine and homogeneous grains of Nb3Sn were studied through a series of experiments [12,13]. The phase transformations of high Jc Nb3Sn strands during reaction heat treatment were studied by performing synchrotron X-ray diffraction (XRD) measurements.
The amount of Sn plays an important role in the superconducting properties of Nb3Sn. The critical temperature sharply increases from approximately 6 K to approximately 18 K when the percentage of Sn is increased from below 20% to above 24.5%. For some superconductors in which Sn concentration is insufficient or some regions with a deficient stoichiometric ratio of Sn, the critical temperature of Nb3Sn may be below 9 K [14]. The upper critical field is also affected by the amount of Sn in accordance with the Werthamer–Helfand–Hohenberg theory. Therefore, increasing the amount of Sn or making the stoichiometry of the Nb3Sn layer more uniform is important to increase the upper critical field and the critical temperature [4]. The addition of Cu can decrease A15 formation temperature and obtain Nb3Sn grains with small sizes by limiting grain growth. Thus, higher grain boundary density can be obtained to improve the bulk pinning force [15]. A typical reaction scheme for high-performance Nb3Sn conductors is increasing the temperature to 483 K and holding this temperature for 100 h, ramping the temperature to 673 K and holding this temperature for 50 h, and then ramping to the final reaction temperature (923–973 K) and holding this temperature for a long period (50–200 h). The heat treatment of the Nb3Sn bulk is equally important because the bulk can be used to fabricate the Nb3Sn superconducting joint, which must be heat treated with Nb3Sn conductors. Reducing the length of the final heat treatment process is significant for Nb3Sn magnet production.
We report a simple and time-saving synthesis technique for Nb3Sn that proceeds through a mechanical alloying method, followed by short-term heat treatment. Nb, Sn, and Cu powders were mixed and alloyed through mechanical alloying, and the precursors were synthesized via high-energy ball milling of the as-blended powders. Then, the precursors were heat treated to form Nb3Sn. However, before Nb3Sn nucleation and growth, other intermetallic phases were synthesized during the large temperature intervals. These phases affected the final reaction in the formation of Nb3Sn. The influence of temperature variation on Nb3Sn reaction was studied, and changes in crystal structure with temperature were measured to study the change in phase transition during the process. After high-energy ball milling at different times, the precursors were prepared and heat treated at various temperatures ranging from 298 K to 1073 K at 10–50 K intervals.

2. Experiment

All original materials used in this work, including Nb, Sn, and Cu, were composed of 99.9% pure powder, which were 5–20 µm in scale. Nb, Sn, and Cu were blended with a molar ratio of 3:1:1. Then, the blended powders were placed in a ball mill tank. The rotational speed of the ball mill was 1725 rpm. The precursors were synthesized after high-energy ball milling. The procedure above was performed in a glove box to prevent the blended powders from oxidizing in a high-temperature and high-pressure environment. The as-blended powders were mechanical alloyed at different times by using a Spex 8000D mixer. The prepared precursors were then heat treated from room temperature to 1073 K. During the process, we studied the changes in the crystal structure of the precursors after the selected temperatures, which was approximately at every 50 K interval from 298 K to 873 K and from 973 K to 1073 K and at every 10 K interval from 903 K to 953 K. The heating rate was 10 K per minute, and sintering holding time was 15 min at each sintering temperature point.
The morphology of the product was recorded using a scanning electron microscope. A transmission electron microscope was used to analyze the crystallinity of the as-prepared samples. The prepared samples were characterized via XRD on an Empyrean X-ray diffractometer. The temperature dependence of magnetizability was measured using a physical property measurement system (PPMS-9).

3. Results and Discussion

Figure 1 shows the scanning electron microscopy (SEM) images of the original samples. Figure 2 shows the SEM images and the energy-dispersive X-ray spectroscopy (EDS) mapping images of the products after the milling of Nb–Sn–Cu powders for 5 h. The images indicated that all the elements were homogeneously dispersed. Patankar [16] reported that brittle Nb with a body-centered cubic structure is likely to get fragmented and get coated onto the surface of ductile Sn. Patankar believed that the depth of XRD was insufficient to penetrate the Nb shell. Thus, Sn was not observed from the XRD result. However, the results of the EDS mapping images showed that Sn and Cu were homogeneously dispersed around Nb and did not exhibit the core–shell structure. This phenomenon may be attributed to the amorphization of Sn with increasing milling time. The original material was several microns in size. After mechanical alloying, the dimensions of the particles decreased. Even nanocrystalline materials were formed due to high-energy ball milling.
Figure 3 depicts the elemental composition of the as-blended particles after mechanical alloying, as illustrated in the inset of the figure. From the results of the selected area with a few microns, Nb, Sn, and Cu were included and mixed in stoichiometric ratio, indicating that a new combination of elements was constructed at the microscopic level. This finding exerted a significant effect on the reaction of Nb with Sn to produce Nb3Sn due to the microscopic stoichiometric ratio and the closer diffusion distance between Nb and Sn in the subsequent heat treatment process.
Figure 4 shows the typical low-magnification and high-resolution transmission electron microscopy (TEM) micrographs of the samples. The selected area electron diffraction pattern shown in the inset of Figure 4a suggested that the products exhibited poor crystallinity after ball milling for 5 h. One study suggested that Sn transformed into the amorphous phase because Nb and Cu were harder than Sn [17]. Some diffraction spots were observed with halo rings in the figure. The Nb grains were confirmed through the high-magnification micrograph in Figure 4b.
The finding was consistent with the XRD result below. Diffraction intensity became weaker with increasing milling time. This result suggested that on the one hand, the crystalline grain size of products became smaller and even reached nanometer scale, broadening the width of diffraction peaks. On the other hand, the high-energy ball milling introduced many defects into the products. The disorder degree of interfacial atoms increased considerably, reducing diffraction intensity. The high-resolution TEM micrograph showed the interplanar crystal spacing of Nb.
Figure 5 presents the XRD patterns of the particles after mechanical alloying at different times. The results showed that the diffraction peaks of Sn and Cu disappeared when milling time reached 1 h compared with the apparent diffraction peaks of the original as-milled particles. The intensity of peaks decreased with increasing milling time. When ball milling time reached 10 h, the primary diffraction peak of Nb evidently broadened compared with the initial stage. High-energy mechanical alloying caused the continuous cold welding and fracturing of different particles. Grain size decreased with increasing milling time. The results showed that the effects of ball milling on the crystallinity of products became worse with increasing milling time. This effect was also observed from the TEM results of the products. Simultaneously, long-range-order parameters decreased with increasing milling time due to the large number of defects and the high amount of grain boundary energy during high-energy mechanical alloying. Both caused the intensity of peaks to decrease and transform into metastable phases.
Figure 6 presents the XRD patterns of the products after different heat treatment temperatures ranging from 298 K to 573 K. These curves represent the crystal structures of the products after each heat treatment at different temperatures. As shown in the figure, the XRD pattern at 298 K presents the measurement result at room temperature, i.e., the crystal structure of the sample after mechanical alloying. Only the Nb diffraction peaks are shown. The Sn and Cu diffraction peaks disappeared after high-energy mechanical alloying.
After heat treatment from room temperature to 573 K, the XRD patterns showed that the products did not change, and no new diffraction peaks appeared. These results indicated that the crystal structures of the products did not change, and no chemical reaction occurred at below 573 K. The melting point of Sn is below 504.9 K, and Sn will liquify at temperatures above 504.9 K. However, the results obtained after the 573 K heat treatment showed that Sn did not change and turn to liquid, signifying that Sn reacted with the other elements to exist in a new form during high-energy mechanical alloying.
The XRD patterns of the particles after heat treatment from 573 K to 773 K are depicted in Figure 7. The shapes of the peak diffraction were nearly identical from 573 K to 723 K. After heat treatment temperature increased to 773 K, small diffraction peaks of Nb3Sn appeared. This finding indicated that reactions that produced Nb3Sn started when the temperature was above 773 K. Simultaneously, the diffraction peaks of Nb began to fade due to the reaction of Nb with Sn or Sn compounds.
As shown in Figure 8, Nb diffraction peaks gradually attenuated after heat treatment from 773 K to 903 K. Nb diffraction peaks nearly disappeared at temperatures above 903 K. In this process, Nb3Sn was produced with increasing temperature. After heat treatment at 903 K, Nb3Sn phase diffraction peaks became highly apparent. Thus, through high-energy mechanical alloying, the as-milled particles should be at least exposed to an environment with a temperature of above 903 K.
To determine the optimum temperature range of heat treatment, temperatures of 903 K, 913 K, and 923 K were selected to study the crystal structure of the products. As shown in Figure 9, the crystal structures were nearly identical at different temperatures, and Nb3Sn and NbO2 diffraction peaks were observed. Nb3Sn diffraction peaks exhibited slightly higher intensity with increasing temperature.
As shown in Figure 10, the samples were heat treated at different temperatures of 923 K, 933 K, and 943 K. The results indicated that the locations of the diffraction peaks did not change, and no new structures appeared. Thus, the temperature of 923 K was selected to obtain the final Nb3Sn product.
The effect of higher heat treatment temperatures was studied to determine whether the crystal structure of the product changed when exposed to higher temperatures. Figure 11 shows the XRD patterns of the products after heat treatment at different temperatures ranging from 953 K to 1073 K. The results showed that the diffraction peaks of Nb3Sn did not change when exposed to temperatures above 973 K, but other impurities, such as Nb2O5, appeared. Thus, as-milled particles should not be utilized at above 973 K.
The key temperature points of Nb3Sn heat treatment were presented to show the change in crystal structure during the heat treatment process. As illustrated in Figure 12, the phase transition process from room temperature to 903 K and then to 1073 K was as follows: the original materials of Nb, Sn, and Cu at 298 K exhibited broadened Nb diffraction peaks at 298 K after mechanical alloying. Then, Nb3Sn and NbO2 diffraction peaks appeared at 903 K. Finally, Nb3Sn, NbO2, and Nb2O3 diffraction peaks appeared at 1073 K. In accordance with the solid-state diffusion reaction, different temperatures correspond to different activation energies of atomic diffusion, affecting the diffusion reaction speed and whether diffusion reaction can occur. In accordance with the result of heat treatment, the as-milled particles should be exposed to temperatures above 903 K. However, when the temperature reached 1073 K, diffraction peaks of impurities appeared.
To maximize the Nb3Sn area, heat treatment time must last for a long time (50–200 h) at high temperature in Nb3Sn formation reaction. The researchers investigated the effects of reaction time (24–150 h) on the superconducting characteristics of Nb3Sn. They determined that a 24 h final reaction reached >90% of the highest obtained Jc [18]. To obtain a short reaction time and good performance, the experimental sample received a final reaction for 24 h at 923 K. With this scheme, we can combine the technologies of Nb3Sn bulk and conductors to fabricate Nb3Sn superconducting joints on magnets in the future.
Figure 13 shows MT curves of the samples that were ball milled for 5 h and heat treated for 24 h and 50 h. The onset Tc of the samples was between 17 K and 18 K. The Tc value was determined by taking the first deviation point from the linearity that signified the transition from the normal state to the superconducting state. As shown in Figure 13, the transitions of all the samples were relatively steep, indicating that the Nb3Sn phase formation qualities of the samples were good. The differences among the prepared samples and between the experiment results and the theoretical value may be attributed to the purity of the products. Critical temperature exhibits a relationship with the concentration of Sn and the stoichiometric ratio of Nb and Sn. After high-energy mechanical alloying, the size of Nb and Sn particles sharply decreased. The distribution of Nb and Sn became more unique, and the solid-state diffusion reaction of Nb and Sn occurred on the basis of a more precise stoichiometric ratio of 3:1. Bc2(0) can be calculated using the following equation based on the Werthamer–Helfand–Hohenberg theory [19]:
B c 2 0 = 0.69 T c B c 2 T | T c
where B c 2 is the upper critical magnetic field, Tc is the critical temperature of a Type II superconductor (Nb3Sn), and 0.69 is the pre-factor constant for the dirty limit [20].
The results led to the conclusion that the   B c 2 0 of the prepared samples did not appear to depend on the heat treatment time of the samples.
Figure 14 shows the field independence of the critical current density Jc of the heat-treated bulk for 24 h. The heat-treated bulk was about 3 mm in length, 2 mm in width, and 1 mm in thickness. The magnetic Jc was calculated from the magnetization loops based on the critical Bean model [21], as follows:
J c = Δ M a 1 a 3 b ,
where a and b (ba) are the width and length of the sample, respectively.
Jc decreased with increasing magnetic fields. The phenomenon of the curve in the low magnetic field is attributed to the flux jump, which typically occurs in Nb3Sn materials at low fields in changing magnetic fields. Flux jumps result in the transition between the superconducting and normal states, which has been readily observed during V–H measurements. Compared with the performance of Nb3Sn in low magnetic field, we are more concerned about the performance of Nb3Sn in a high magnetic field, because the Nb3Sn material will contribute more to the high magnetic field region. As shown in the figure, the magnetic Jc was approximately 30,000 A/cm2 under 8 T. Considering the limit of the magnetic field in the PPMS instrument (±9 T), the Jc value under 8 T was used to evaluate the performance of Nb3Sn. The technology can be used for fabricating Nb3Sn joints in Nb3Sn magnets. The magnetic intensity in the region of Nb3Sn joints is lower than that in the center of the magnet. The result can help magnet designers place the Nb3Sn joint on the location where the joint has no problem to work under such a magnetic field.

4. Conclusions

This study presented the process of Nb3Sn phase transitions during heat treatment. Nb3Sn was prepared by heat treating the precursors, which were mechanically alloyed first. The size and crystal structure of the products changed after mechanical alloying for some time. The TEM results indicated that the products exhibited poor crystallinity after ball milling, and this finding was consistent with the XRD results. In accordance with the XRD analysis, the diffraction peaks of Cu and Sn became weaker with increasing ball milling time, and only Nb diffraction peaks were observed after 3 h. The precursors were heat treated from room temperature to 1073 K, and simultaneously, the crystal structures were monitored after exposure to different temperatures. When the temperature was above 773 K during heat treatment, Nb3Sn began to appear. When the temperature was above 973 K, some impurities, such as Nb2O5, appeared. Thus, the heat treatment temperature of the as-milled particles should not be above 973 K. The critical temperatures of the samples that were heat treated at different times were between 17 K and 18 K. The magnetic critical current density of the sample versus the applied magnetic field at 4.2 K illustrated that magnetic Jc was approximately 30,000 A/cm2 under 8 T. The result can be applied to the fabrication of Nb3Sn superconducting joints in the future and help magnet designers place the joint in an appropriate location where the joint will not be affected by magnetic intensity.

Author Contributions

Conceptualization, W.S. and S.C.; methodology, W.S.; software, W.S.; validation, W.S.; formal analysis, W.S.; investigation, W.S.; resources, W.S. and S.C.; data curation, W.S.; writing—original draft preparation, W.S.; supervision, S.C.; project administration, S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Matthias, B.; Geballe, T.; Geller, S.; Corenzwit, E. Superconductivity of Nb3Sn. Phys. Rev. 1954, 95, 1435. [Google Scholar] [CrossRef]
  2. Baig, T.; Yao, Z.; Doll, D.; Tomsic, M.; Martens, M. Conduction cooled magnet design for 1.5 T, 3.0 T and 7.0 T MRI systems. Supercond. Sci. Technol. 2014, 27, 125012. [Google Scholar] [CrossRef]
  3. Iwasa, Y.; Bascuñán, J.; Hahn, S.; Park, D.K. Solid-Cryogen Cooling Technique for Superconducting Magnets of NMR and MRI. Phys. Procedia 2012, 36, 1348–1353. [Google Scholar] [CrossRef] [Green Version]
  4. Godeke, A.; Hellman, F.; Kate, H.H.J.T.; Mentink, M.G.T. Fundamental origin of the large impact of strain on superconducting Nb3Sn. Supercond. Sci. Technol. 2018, 31, 105011. [Google Scholar] [CrossRef] [Green Version]
  5. Iwaki, G.; Sato, J.; Inaba, S.; Kikuchi, K. Development of bronze-processed Nb3Sn superconducting wires for high field magnets. IEEE Trans. Appl. Supercond. 2002, 12, 1045–1048. [Google Scholar] [CrossRef]
  6. Miyazaki, T.; Kato, H.; Hase, T.; Hamada, M.; Murakami, Y.; Itoh, K.; Kiyoshi, T.; Wada, H. Development of High Sn Content Bronze Processed Nb3Sn Superconducting Wire for High Field Magnets. IEEE Trans. Appl. Supercond. 2004, 14, 975–978. [Google Scholar] [CrossRef]
  7. Gregory, E.; Pyon, T. Internal tin Nb3Sn conductor development for high energy physics applications. Aip Conf. Proc. 2002, 614, 958–967. [Google Scholar]
  8. Lu, X.F.; Hampshire, D.P. The field, temperature and strain dependence of the critical current density of a powder-in-tube Nb3Sn superconducting strand. Supercond. Sci. Technol. 2010, 23, 025002. [Google Scholar] [CrossRef] [Green Version]
  9. Di Michiel, M.; Scheuerlein, C. Phase transformations during the reaction heat treatment of powder-in-tube Nb3Sn superconductors. Supercond. Sci. Technol. 2007, 20, 55–58. [Google Scholar] [CrossRef] [Green Version]
  10. Xu, X. A review and prospects for Nb3Sn superconductor development. Supercond. Sci. Technol. 2017, 30, 093001. [Google Scholar] [CrossRef] [Green Version]
  11. Di Michiel, M.; Scheuerlein, C. Phase transformations during the reaction heat treatment of Nb3Sn superconductors. J. Phys. Conf. 2010, 234, 022032. [Google Scholar] [CrossRef] [Green Version]
  12. Oh, S.; Park, S.H.; Lee, C.; Choi, H.; Kim, K. Heat treatment effect on the strain dependence of the critical current for an internal-tin processed Nb3Sn strand. Physical C 2010, 470, 129–133. [Google Scholar] [CrossRef]
  13. Zhang, K.; Zhang, P.X.; Shi, Y.G.; Liu, J.W.; Gao, H.X.; Jia, J.J.; Guo, J.H.; Li, J.F.; Liu, X.H.; Feng, Y. An Investigation Into the Heat Treatment Tolerance of WST Nb3Sn Strands Produced for Massive Fusion Coils. IEEE Trans. Appl. Supercond. 2015, 25, 1–6. [Google Scholar] [CrossRef]
  14. Lee, J.; Posen, S.; Mao, Z.; Trenikhina, Y.; He, K.; Hall, D.L.; Liepe, M.; Seidman, D.N. Atomic scale analysis of Nb3Sn on Nb prepared by a vapor-diffusion process for superconducting radiofrequency cavity applications. arXiv 2019, 32, 024001. [Google Scholar]
  15. Godeke, A. A review of the properties of Nb3Sn and their variation with A15 composition, morphology and strain state. Supercond. Sci. Technol. 2006, 19, 68–80. [Google Scholar] [CrossRef] [Green Version]
  16. Patankar, S.N.; Froes, F.H. Formation of Nb3Sn Using Mechanically Alloyed Nb—Sn Powder. Solid State Sci. 2004, 6, 887–890. [Google Scholar] [CrossRef]
  17. Menary, M.A.; Menary, M.A. The Effect of Mechanical Alloying on the Structural and Superconducting Properties of Nb3Sn. 2005, Volume 27, pp. 3–7. Available online: https://www.mendeley.com/catalogue/24a95fbb-8978-3aaa-991e-66b759e6088e/ (accessed on 23 February 2023).
  18. Ghosh, A.K.; Sperry, E.A.; Cooley, L.D.; Moodenbaugh, A.M.; Sabatini, R.L.; Wright, J.L. Dynamic stability threshold in high-performance internal-tin Nb3Sn superconductors for high field magnets. Supercond. Sci. Technol. 2004, 18, 5–8. [Google Scholar] [CrossRef] [Green Version]
  19. Helfand, E.; Werthamer, N.R. Temperature and Purity Dependence of the Superconducting Critical Field, Hc2. Phys. Rev. Lett. 1964, 13, 686–688. [Google Scholar] [CrossRef]
  20. Rekaby, M.; Awad, R.; Abou-Aly, A.I.; Yousry, M. AC Magnetic Susceptibility of Y3Ba5Cu8O18 Substituted by Nd3+ and Ca2+ Ions. J. Supercond. Nov. Magn. 2019, 32, 3483–3494. [Google Scholar] [CrossRef]
  21. Bean, C.P. Magnetization of High-Field Superconductors. Rev. Mod. Phys. 1964, 36, 886–901. [Google Scholar] [CrossRef]
Figure 1. SEM images of original samples: (a) Nb, (b) Sn, and (c) Cu.
Figure 1. SEM images of original samples: (a) Nb, (b) Sn, and (c) Cu.
Crystals 13 00660 g001
Figure 2. SEM image and EDS mapping of the samples after the milling of the Nb–Sn–Cu powder for 5 h.
Figure 2. SEM image and EDS mapping of the samples after the milling of the Nb–Sn–Cu powder for 5 h.
Crystals 13 00660 g002
Figure 3. EDS analysis of the sample after the milling of the Nb–Sn–Cu powder for 5 h.
Figure 3. EDS analysis of the sample after the milling of the Nb–Sn–Cu powder for 5 h.
Crystals 13 00660 g003
Figure 4. TEM micrograph of the sample: (a) low-magnification micrograph and (b) high-magnification micrograph.
Figure 4. TEM micrograph of the sample: (a) low-magnification micrograph and (b) high-magnification micrograph.
Crystals 13 00660 g004
Figure 5. XRD patterns of the particles after mechanical alloying at different times.
Figure 5. XRD patterns of the particles after mechanical alloying at different times.
Crystals 13 00660 g005
Figure 6. XRD patterns of products after heat treatment at different temperatures (from 298 K to 573 K).
Figure 6. XRD patterns of products after heat treatment at different temperatures (from 298 K to 573 K).
Crystals 13 00660 g006
Figure 7. XRD patterns of the products after heat treatment at different temperatures (from 573 K to 773 K).
Figure 7. XRD patterns of the products after heat treatment at different temperatures (from 573 K to 773 K).
Crystals 13 00660 g007
Figure 8. XRD patterns of the products after heat treatment at different temperatures (from 773 K to 903 K).
Figure 8. XRD patterns of the products after heat treatment at different temperatures (from 773 K to 903 K).
Crystals 13 00660 g008
Figure 9. XRD patterns of the products after heat treatment at different temperatures (from 903 K to 923 K).
Figure 9. XRD patterns of the products after heat treatment at different temperatures (from 903 K to 923 K).
Crystals 13 00660 g009
Figure 10. XRD patterns of the products after heat treatment at different temperatures (from 923 K to 943 K).
Figure 10. XRD patterns of the products after heat treatment at different temperatures (from 923 K to 943 K).
Crystals 13 00660 g010
Figure 11. XRD patterns of the products after heat treatment at different temperatures (from 953 K to 1073 K).
Figure 11. XRD patterns of the products after heat treatment at different temperatures (from 953 K to 1073 K).
Crystals 13 00660 g011
Figure 12. XRD patterns of the products after heat treatment at different temperatures (from 298 K to 1073 K).
Figure 12. XRD patterns of the products after heat treatment at different temperatures (from 298 K to 1073 K).
Crystals 13 00660 g012
Figure 13. MT curve of the samples that were heat treated for 24 h and 50 h.
Figure 13. MT curve of the samples that were heat treated for 24 h and 50 h.
Crystals 13 00660 g013
Figure 14. Magnetic critical current density of the sample vs. applied magnetic field at 4.2 K.
Figure 14. Magnetic critical current density of the sample vs. applied magnetic field at 4.2 K.
Crystals 13 00660 g014
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sun, W.; Chen, S. Phase Transition of Nb3Sn during the Heat Treatment of Precursors after Mechanical Alloying. Crystals 2023, 13, 660. https://doi.org/10.3390/cryst13040660

AMA Style

Sun W, Chen S. Phase Transition of Nb3Sn during the Heat Treatment of Precursors after Mechanical Alloying. Crystals. 2023; 13(4):660. https://doi.org/10.3390/cryst13040660

Chicago/Turabian Style

Sun, Wanshuo, and Shunzhong Chen. 2023. "Phase Transition of Nb3Sn during the Heat Treatment of Precursors after Mechanical Alloying" Crystals 13, no. 4: 660. https://doi.org/10.3390/cryst13040660

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