Next Article in Journal / Special Issue
The Influence of Preparation Temperature on the Different Facets of Bulk MgB2 Superconductors
Previous Article in Journal
Polymeric and Paper-Based Lab-on-a-Chip Devices in Food Safety: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Performance of the Two-Seeded GdBCO Superconductor Bulk with the Buffer by the Modified TSMG Method

1
College of Mathematics and Physics, Shanghai University of Electric Power, Shanghai 201306, China
2
Shanghai Key Laboratory of High Temperature Superconductors, Shanghai University, Shanghai 200444, China
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(5), 987; https://doi.org/10.3390/mi14050987
Submission received: 28 March 2023 / Revised: 24 April 2023 / Accepted: 26 April 2023 / Published: 30 April 2023

Abstract

:
The multiseeding technique is a method to grow large-sized REBa2Cu3O7−δ (REBCO, where RE is a rare earth element) high temperature superconducting bulks. However, due to the existence of grain boundaries between seed crystals, the superconducting properties of bulks are not always better than those of single grain bulks. In order to improve the superconducting properties caused by grain boundaries, we introduced buffer layers with a diameter of 6 mm in the growth of GdBCO bulks. Using the modified top-seeded melt texture growth method (TSMG), that is, YBa2Cu3O7−δ (Y123) as the liquid phase source, two GdBCO superconducting bulks with buffer layers with a diameter of 25 mm and a thickness of 12 mm were successfully prepared. The seed crystal arrangement of two GdBCO bulks with a distance of 12 mm were (100/100) and (110/110), respectively. The trapped field of the GdBCO superconductor bulks exhibited two peaks. The maximum peaks of superconductor bulk SA (100/100) were 0.30 T and 0.23 T, and the maximum peaks of superconductor bulk SB (110/110) were 0.35 T and 0.29 T. The critical transition temperature remained between 94 K and 96 K, with superior superconducting properties. The maximum JC, self-field of SA appeared in specimen b5, which was 4.5 × 104 A/cm2. Compared with SA, the JC value of SB had obvious advantages in a low magnetic field, medium magnetic field and high magnetic field. The maximum JC, self-field value appeared in specimen b2, which was 4.65 × 104 A/cm2. At the same time, it showed an obvious second peak effect, which was attributed to Gd/Ba substitution. Liquid phase source Y123 increased the concentration of the Gd solute dissolved from Gd211 particles, reduced the size of Gd211 particles and optimized JC. For SA and SB under the joint action of the buffer and the Y123 liquid source, except for the contribution of Gd211 particles to be the magnetic flux pinning center with the improvement of JC, the pores also played a positive role in improving the local JC. More residual melts and impurity phases were observed in SA than in SB, which had a negative impact on the superconducting properties. Thus, SB exhibited a better trapped field and JC.

1. Introduction

There have been decades of research history since the discovery of (RE)Ba2Cu3O7−δ ((RE)BCO, where RE is a rare earth element) high-temperature superconductors. RE(BCO) high-temperature superconductors have broad application prospects and huge application potential in high-tech fields, such as superconducting maglev trains, flywheel energy storage systems, and superconducting bearings [1,2,3,4,5]. The development of RE(BCO) high-temperature superconductors is gradually becoming mature, but in its research process, there are also many problems that need to be solved. The single-grain high-temperature superconductor bulks manufactured by the top-seeded melt growth process (TSMG) have been studied [6,7,8] and are usually grown as parallelepiped with (100), (010), and (001) crystal planes [9]. However, the superconducting performance is unable to reach the expected results because of its slow growth rate and microstructure defects. Traditional TSMG technology has many defects due to the outflow of the liquid phase in the heating process, the generation of oxygen, and unstable factors in the powder pressing process, such as the obvious shrinkage of the final sample, the existence of a large number of pores, and the region without the second phase particles in the microstructure [10,11,12]. In order to solve these problems, many new research methods have been proposed sequentially. Infiltration growth (IG) methods can supply enough liquid sources to reduce the shrinkage of the final bulks, compared with the TSMG method [13]. The bulks grown by Melt-Powder-Melt-Growth (MPMG) have fewer weak links, bigger superconductor grains, smaller RE211 particles, and less pores [14,15].
Buffer-assisted technology is a recognized effective method to improve the performance of high-temperature superconductor bulks [16,17,18,19,20]. By optimizing the composition and the aspect ratio of the buffer layer, Namburi et al. [21] improved the problems related to interfacial stress and seed contamination, and then YBCO bulk superconductors with a diameter of 25~35 mm were successfully prepared. As recorded in Ref. [20], first, the use of the buffer greatly improves the seeding success rate, which can prevent mutual pollution between the seed crystal and precursor block and can absorb the macro cracks caused by lattice mismatch so as to improve the superconducting properties and thermal stability of the seed crystal. Second, the buffer can be used as a large secondary seed to induce the homogeneous epitaxial growth of the main particles. Increasing the size of the c-grown region and the crystallinity of the grown crystal can also homogenize the distribution of the Gd211 particles and improve the critical current density. With the simultaneous introduction of the buffer and the rich liquid phase, GdBCO-Ag bulk samples with a diameter of 48 mm have been successfully prepared [22].
Multiseeding technology is expected to prepare large-size REBCO bulk samples [23,24,25,26]. However, multiseeding technology will reduce the critical current density characteristics of superconductor bulks due to the existence of a residual impurity phase at the grain boundary between seed crystals. Therefore, only when the grain boundaries are eliminated are large overcurrent loops generated, resulting in large magnetic moments and trap magnetic fields [21]. Samples with seed crystals placed in (110/110) mode have relatively clean grain boundaries [27]. In addition, Shi et al. used bridge seeding technology to achieve the best alignment in the preparation of multi-seeded samples. They found that the seed (110/110) arrangement has more advantages than the seed (100/100) arrangement because this arrangement can form clean grain boundaries [28,29]. As documented in reference [30], samples with different arrangements have different growth patterns at grain boundaries. If two grains impinging each other are planar parallel (100/100), each grain grows in the opposite direction and then meets and stops at the grain boundary, where the residual melt cannot continue. On the other hand, if two grains impacting each other are not parallel in plane (110/110), and the grains meet and continue to grow along the grain boundary, there will be no residual melt.
Compared with other light rare earth elements, the GdBCO superconducting bulk has the most stable performance and the lowest Gd/Ba substitution effect, and the element substitution to a lesser extent can deepen the second peak effect and improve the critical current density. In this paper, the superconducting properties and microstructure of GdBCO high temperature superconductor bulks prepared by the combination of two- seeded technology in different arrangement and buffer technology are studied, which is a new test to grow the high performance of the superconductor bulks. The superconductor material prepared in this paper is a GdBCO bulk. Double NdBCO seeds and double buffer GdBCO superconductor bulks with a diameter of 25 mm were prepared by the modified top-seeded melt growth process (TSMG), and the YBa2Cu3O7−δ (Y123) pellet was introduced as the liquid phase source to avoid the shrinkage problem caused by liquid phase loss [31,32]. The effects of different arrangement of seed crystals on the properties of the bulks were studied. The superconducting properties were analyzed by the trapped field, critical current density (JC), critical transition temperature (TC) and microstructure.

2. Experimental

According to GdBa2Cu3O7−δ + 40 mol% Gd2BaCuO5 + 10 wt%Ag2O + 0.5 wt%Pt, commercial GdBa2Cu3O7−δ (Gd123, 99.9%) powder was mixed with Gd2BaCuO5 (Gd211, 99.9%) powder in the molar ratio of 1:0.4, and 10 wt% Ag2O and 0.5 wt% Pt were added to improve the mechanical strength and refine Gd211 particles. They were ground evenly with a mortar and then pressed into a cylindrical pellet with a diameter of 25 mm and a thickness of 12 mm as the precursor. Commercial Y123 (99.9%) powder was used in this experiment as a liquid phase source; it was pressed into a cylindrical pellet with a diameter of 25 mm and a thickness of 3 mm. The chemical compositions of buffer pellets were similar to those of precursors, except for not adding Ag2O and Pt powder. After being fully ground, they were pressed into cylindrical pellets with a diameter of 6 mm and a thickness of 3 mm. Seed crystals with a size of 2 × 2 × 0.5 mm3 NdBCO were used to guide the growth.
The pressed pellets were assembled in order from top to bottom: NdBCO seed crystals, buffers, precursor, Y123 liquid source, Y2O3 and aluminum oxide plate. A Y2O3 cylindrical pellet with equal diameter was inserted between the Al2O3 plate and the Y123 pellet to prevent a reaction between them. As shown in Figure 1, two NdBCO seed crystals were arranged in two different ways: (100/100), and (110/110), named SA and SB. The distance between the two seed crystals (the linear distance between the center of the seed crystal) was 12 mm.
The assembled pellet was placed into a box furnace for sintering, and the temperature was raised to 1080 °C at a rate of 50 °C/h, held for 1 h, quickly cooled to 1009 °C, and then slowly cooled to 979 °C at a rate of 0.3 °C/h. Finally, it was cooled to room temperature with the furnace, as shown in Figure 2. In order to obtain the orthogonal superconducting phase structure, annealing treatment was carried out in a flowing high-purity oxygen atmosphere. SA and SB were heated to 450 °C within 5 h, kept for 40 h, and then cooled to 350 °C within 40 h and 300 °C within 30 h, respectively. In the end, they were naturally cooled to room temperature. Then, two GdBCO bulks with high performance were fabricated successfully.
The top and bottom surfaces of the bulks were polished to measure the trapped field. Under an external magnetic field of 1.5 T and liquid nitrogen environment for 30 min, the external magnetic field was removed, and the trapped magnetic field curve was tested by a Hall probe sensor with a distance of 2 mm from the surface of the bulks. In order to study the effect of the buffer layer on the performance of the bulk, based on the growth patterns of two bulk materials, namely along the direction parallel to the edge of the seed crystal, we cut five specimens of b1, b2, b3, b4 and b5 from SA and SB under the position of the seed crystal. The surface of the bulks was polished about 1 mm, and the b1~b5 specimens with a size of 2 mm × 2 mm × 1 mm were took from the upper surface to the bottom, as shown in Figure 3c. The physical property measurement system (PPMS) was used to measure the DC magnetization of specimens cut at different positions of the bulks. The critical current density (JC) was calculated according to the extended Bean critical state model. The microstructure of the bulks was observed by scanning electron microscopy (SEM).

3. Results and Discussion

The macromorphology of SA and SB is shown in Figure 3. The growth facet line of the seed crystal is marked with a blue dotted line in the picture. It can be seen that the seeds were intact, and there was no melting phenomenon. There was an obvious drift phenomenon between the two seed crystals of SA, and the two seed crystals of SB were arranged relatively orderly, but there was still a slight drift phenomenon (indicated by the red arrow). From the top views, we can clearly see the grain boundary. The appearance of grain boundaries is inevitable in the growth process, so we need to reduce the negative impact of grain boundaries on the superconductor bulk. Many factors affect the grain boundary. Theoretically, the factors affecting grain boundary depth are the distance between seed crystals. The arrangement of the seed crystals and the buffer layer and the size of the buffer layer are also important factors when using the buffer layer. The depth of the grain boundary increases with the increase of the distance between seed crystals for the growth of the multiseeded bulk. The depth of the grain boundary of multiseeded bulks is mainly determined by the angle [33] formed by the top surface and the boundary of the a and c growth region. The way to change this angle is to increase or reduce the distance between the seed crystals. The larger the angle, the deeper the grain boundary is formed, indicating that the distance between the seed crystals is farther. The smaller the angle, the shallower the grain boundary formed, indicating that the distance between the seed crystals is closer. The seed crystal and the buffer layer can effectively induce growth as a large-size seed when the buffer layer is added, reduce the effective length between the seed crystals significantly, reduce the a-sector growth area and increase the c-sector growth area. Then, the angle formed between the top surface and the boundary of the a-sector and c-sector growth area is reduced so as to decrease the depth of the grain boundary, which improves the superconducting properties [17,34,35,36].
Figure 4 shows the distribution of trapped magnetic flux density on the top surface of two bulks measured in a liquid nitrogen environment (3D map and 2D map). It can be seen from the picture that the magnetic flux distribution of the two bulks is conical and shows two peaks. The two peaks of SA were 0.30 T and 0.23 T, and the two peaks of SB were 0.35 T and 0.29 T, respectively. SB had a smaller difference between the two peaks and showed a higher peak compared with SA. It can be observed in (a) and (b) of Figure 4 that the two growth regions were not symmetrical, and the nucleation particles guided by one seed crystal were significantly better than the other. This phenomenon may be caused by the different nucleation times of two seed crystals, the different temperatures at the different positions of the seed crystals, or the asymmetric arrangement of two-seeded crystals, which should occur randomly [28].
The symmetry of the arrangement of two-seeded crystals of SA (100/100) was not high, which led to poor connectivity in the grains and wrong crystal orientation between the growth regions to form polluted grain boundaries. The accumulation of non-superconducting phases is the obstruction of the flow of the superconducting current in the current circuit, which leads to the poor performance of the trapped field [20,37,38]. In contrast, the seed crystal arrangement symmetry of SB was higher. In the (110/110) seed crystal arrangement, the grains met at the grain boundary along the growth direction, which continued to grow to form a relatively clean grain boundary. Therefore, both the value of the peak and the difference between the two peaks were better, which suggests that the trapped field of SB was better than SA. Therefore, we infer that the trapped field distribution of the GdBCO superconductor bulk guided by a multiseed crystal is highly related to the symmetry of the seed crystal and the existence of an impurity phase at the boundary. The higher the symmetry of the seed crystal, the cleaner the grain boundary, and the better the trapped field performance.
Figure 5 shows the distribution of trapped magnetic flux density on the bottom surface of GdBCO superconductor bulks measured in a liquid nitrogen environment (3D map and 2D map). The magnetic flux distribution of the two bulks was conical and showed two peaks. We also found that the two peaks of SA were 0.29 T and 0.27 T, and the two peaks of SB were 0.29 T and 0.33 T. Compared with the top surface, the difference between the two peaks of each bulk on the bottom surface was smaller. This should be a positive effect brought about by the introduction of Y123 as a liquid source in the experiment, which has a significant effect on improving the uniformity of trapped field distribution [39]. Inserting the Y123 phase into the bottom of the precursor not only successfully solves the problem of liquid phase loss in the growth process but also accumulates a sufficient liquid phase at the growth front and plays a great role in leading to the more uninform of the distribution of Gd211 particles in the Gd123 matrix. As we all know, the distribution of Gd211 particles in the Gd123 matrix is an important index to judge the performance of the bulks because they are not only an important source of the flux pinning center, but they also reduce the residue of the unreacted liquid phase to the greatest extent. In addition, the introduction of the Y123 liquid phase source also expands the growth region of c orientation and inhibits the formation of a high angle grain boundary, which plays a positive role in the formation of an excellent trapped field [40].
Figure 6 shows the temperature dependence of the magnetization of specimens b1, b2, b3, b4 and b5 cut below the seed crystals of SA and SB. The curves in Figure 6a,b show that the values of initial transition temperature TC, onset is between 94~96 K with the superior superconducting properties. On the whole, the transition width of SA changes greatly, which is not as sharp as SB. The distribution of TC, onset of SA was relatively concentrated, showing a large transition width at position b1, while it was relatively narrow at the other positions. The distribution of TC, onset of SB was relatively scattered, but the transition width of all specimens was very sharp except b1, and the value of transition width was about 0.5 K. Among them, the TC, onset of position b1 was the lowest, and the TC, onset of position b3 was the highest. This suggests that the transition width of TC in specimen b1 near the seed crystals slightly increases due to the substitution of Gd/Ba ions [25,31]. We infer that the addition of two independent buffer layers has no adverse effect on the superconducting properties of GdBCO bulks.
Figure 7 shows the critical current density curve of specimens b1, b2, b3, b4 and b5 under SA and SB seed crystals. For convenience in the analysis, the self-field JC values of the specimens at different positions are plotted in Figure 8. The JC curve results of SA showed that under the applied magnetic field, the JC, self-field value of the specimens had different trends. Specimen b5 at the bottom had the highest JC, self-field under the low field and was lower than b3 under the middle field. Its highest JC, self-field value was 4.5 × 104 A/cm2. The JC curve results of SB showed that under the applied magnetic field, the JC, self-field of the specimens changed with the increase in the distance from the seed crystal in the c-axis direction, increased from position b1 to b3, and was higher than the value of SA at the corresponding position, but the values of positions b4 and b5 were slightly lower than that of SA. The maximum JC, self-field value of specimen b2 was 4.56 × 104 A/cm2, which is comparable to the results of GdBCO superconductor bulks [25,29]. SB maintained absolute advantages in low magnetic field, medium magnetic field and high magnetic field. The JC value of the bottom specimens of the bulks was very high, which may have been due to the introduction of a Y123- rich liquid phase, which plays a positive role in filling the defects in the microstructure completely and improving the uniformity of superconducting phase particle distribution [39], consistent with the results in trapped fields.
The applied magnetic field started from 5 T; the JC value of the curve in SA was almost zero, while SB was much better. Under low magnetic field and medium magnetic field, SB showed a higher JC, self-field value and more neat curve trend. Moreover, there was a second peak effect near the seed crystal. This effect shows the superiority in the performance of the trapped field. The TC at the corresponding position decreased, and this effect decreased with the increase in the distance from the seed crystal, which is attributed to Gd/Ba substitution [41,42]. The additional Y123 liquid source increases the concentration of Gd solute decomposed from Gd211 particles to strengthen the Gd/Ba substitution effect and improve JC. Because there is a temperature gradient in the process of bulk growth, especially at a higher temperature near the seed crystal, the Gd/Ba substitution effect is more serious, which shows the most obvious second peak phenomenon [43].
Figure 9 is a set of scanning electron microscope (SEM) photographs of specimens b1, b3 and b5 of SA (100/100) and SB (110/110) at 500 times magnification. The pore distribution in the pictures showed typical characteristics, which changed with the distance from the seed crystal. The existence of these pores was mainly due to the precipitation of the gas in the melting process, which were trapped in the melt and forms circular pores. The distribution trend of the pores of two bulks in b1, b3 and b5 was consistent. The number and size of pores at b1 were small, and the number and size of pores at b3 and b5 were large. Compared with that in SA, the grain boundary was narrow and the amount of residual melt at the grain boundary was small in SB specimens, which was related to the high trapped field in SB, as shown in Figure 4 and Figure 5. The pore distribution of SA showed that the number of pores in the bottom was higher than that in the middle, and the middle was higher than that in the top. The pore distribution of SB showed that the number in the middle was higher than that in the bottom, and the number in the bottom was higher than that in the top position. Overall, the number of pores in SA was significantly higher than that in SB.
Figure 10 shows scanning electron microscope photographs of specimens b1, b3 and b5 of SA (100/100) and SB (110/110) at 5000 times magnification. In order to further analyze the Gd211 particles distributed in these GdBCO bulks, the size of the Gd211 particles shown in Figure 10 was determined by the software of Nano measure 1.2, as shown in Figure 11. The average size values of the Gd211 particles (AVG) were obtained from the peak position of each fitted curve. The size of Gd211 particles near the seed crystal was the largest and changed with the increase in the distance from the seed crystal. As shown by comparing the average size of Gd211 particles, the average value of b3 belonging to SB was the smallest, followed by b5 of SA, which had a maximum JC, as shown in Figure 7. This shows that the smaller the size of Gd211 particles, the larger the JC value. Generally, a low-concentration Gd211 region will be formed near the seed crystal. However, in the SEM image of b1, a large number of evenly distributed Gd211 particles was observed, which is a positive effect brought about by the use of a buffer layer [16]. According to the pushing/trapping theory [44], the number of Gd211 distributed at different positions increases with the increase of the distance, and the effect of the particle refinement is better. Due to the addition of the Y123-rich liquid phase, the uniformity of the second phase particles is improved to increase the superconducting performance at this position. From the microstructure (SEM) observation of the bulks, the small change in the volume fraction of Gd211 particles along the c-axis is not effective in explaining the change trend of JC. Then, there are more pores in the area with a high JC value, which suggests that pores of a suitable size may form an effective flux pinning center to improve the local JC. For SA and SB under the joint action of buffer and Y123 liquid source, except for the contribution of Gd211 particles to be the magnetic flux pinning center with the improvement of the JC, porosity also affects performance. Figure 7 shows that the JC of b1 was the lowest, while b3 and b5 increased significantly. For SA, b5 > b3 > b1, and for SB, b3 > b5 > b1. The pore distribution of these specimens was consistent with the change trend of JC, and the pores played a positive role in improving the superconducting properties.
The microstructure observation of SA showed that they were in the solidified liquid phase, and there were impurities that hindered the flow of the superconducting current and reduced the superconducting performance [45]. Moreover, there were many large-sized Gd211 particles that could not be used as an effective flux pinning center to reduce JC. The microstructure observation results of SB showed that the solidified liquid phase and impurities were significantly less than that of SA. It was clean and flat under a scanning electron microscope (SEM), indicating that SB possessed higher and superior superconducting performance, as demonstrated by the exhibition of the trapped field and JC.

4. Conclusions

The buffer layer with a diameter of 6 mm and YBa2Cu3O7−δ (Y123) liquid source were introduced to successfully prepare two GdBCO superconductor bulks with two NdBCO seed crystal arrangements of (100/100) and (110/110) in the distance of 12 mm and then their superconducting properties were carefully analyzed. The diameter and thickness of the precursor of two GdBCO superconductor bulks were 25 mm and 12 mm, respectively. The maximum peaks of both GdBCO superconductor bulks appeared on the top surface, which were 0.30 T in SA and 0.35 T in SB. The use of the buffer layer reduced the depth of the grain boundary and improved the trapped field properties of bulks as a whole. The difference in the peak values of the two bulks was related to the arrangement of seed crystals. The low peak value of SA was due to the existence of a residual melt and impurity phase at the grain boundary, which hindered the movement of superconducting current and led to a poor trapped field. The grain boundary of SB was relatively clean, so its trapped field was better. The magnetic flux distribution on the bottom surface showed similar and high characteristics due to the introduction of the rich Y123 liquid phase. The rich Y123 liquid phase can promote the uniform distribution of Gd2BaCuO5 (Gd211) particles, expand the growth region of c orientation and inhibit the formation of high angle grain boundaries. The critical transition temperature of the bulks was 94–96 K, with good superconducting properties. In the JC0H curve, SB had obvious advantages in a low magnetic field, medium magnetic field and high magnetic field. An obvious second peak effect was observed in the specimens near the seed crystal position, which made the JC value increase and the TC at the corresponding position decrease, and this effect decreased with the increase in the distance from the seed crystal position due to the effect of Gd/Ba substitution. Moreover, the Y123 liquid source increased the concentration of Gd solute dissolved from Gd211 particles and strengthened the Gd/Ba substitution effect. Based on the distribution of Gd211 and pores in superconductor bulks, the smaller the size of the Gd211 particles, the higher the JC of the bulk. The distribution of pores was consistent with the trend of the JC. This suggests that pores also have a positive effect on the superconducting properties of the superconductor bulks. More residual melts and impurity phases existed in SA with the two-seeded arrangement of (100/100), resulting in slightly lower superconducting properties compared with that of SB with the two-seeded arrangement of (110/110). The experimental data will be important for the preparation of large-sized superconductor bulks.

Author Contributions

Conceptualization, Y.Z. (Yufeng Zhang); draft preparation, Y.Z. (Yufeng Zhang) and C.L.; precursor synthesis, Z.L., P.Z. and Y.Z. (Yan Zhang); experiments, C.L., S.S. and G.R.; data analysis, C.L., Z.L. and J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 11004129), the Scientific Research Starting Foundation for the Returned Overseas Chinese Scholars, the Ministry of Education of China (SRF for RCS, SEM), the Innovation Program of Shanghai Municipal Education Commission, China (Grant No. 11YZ197), and the Opening Project of Shanghai Key Laboratory of High Temperature Superconductors (Grand No. 19DZ2270500).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brandt, E.H. Rigid levitation and suspension of high-temperature superconductors by magnets. Am. J. Phys. 1990, 58, 43–49. [Google Scholar] [CrossRef]
  2. Strasik, M.; Hull, J.R.; Mittleider, J.A.; Gonder, J.F.; Johnson, P.E.; McCrary, K.E.; McIver, C.R. An overview of Boeing flywheel energy storage systems with high-temperature superconducting bearings. Supercond. Sci. Technol. 2010, 23, 034021. [Google Scholar] [CrossRef]
  3. Werfel, F.N.; Floegel-Delor, U.; Riedel, T.; Goebel, B.; Rothfeld, R.; Schirrmeister, P.; Wippich, D. Large-scale HTS bulks for magnetic application. Phys. C Supercond. 2013, 484, 6–11. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Zhou, D.; Ida, T.; Miki, M.; Izumi, M. Melt-growth bulk superconductors and application to an axial-gap-type rotating machine. Supercond. Sci. Technol. 2016, 29, 044005. [Google Scholar] [CrossRef]
  5. Peczkowski, P.; Szterner, P.; Jaegermann, Z.; Kowalik, M.; Zalecki, R.; Woch, W.M. Effects of Forming Pressure on Physicochemical Properties of YBCO. J. Supercond. Nov. Magn. 2018, 31, 2719–2732. [Google Scholar] [CrossRef]
  6. Nariki, S.; Sakai, N.; Murakami, M. Melt-processed Gd–Ba–Cu–O superconductor with trapped field of 3 T at 77 K. Supercond. Sci. Technol. 2005, 18, S126–S130. [Google Scholar] [CrossRef]
  7. Cheng, L.; Li, T.; Yan, S.; Sun, L.; Yao, X.; Puzniak, R. Large Size and High Performance of a Gd-Ba-Cu-O Bulk Superconductor Grown Using New Approaches. J. Am. Ceram. Soc. 2011, 94, 3139–3143. [Google Scholar] [CrossRef]
  8. Shi, Y.; Babu, N.H.; Iida, K.; Yeoh, W.; Dennis, A.; Pathak, S.; Cardwell, D. Batch-processed GdBCO–Ag bulk superconductors fabricated using generic seeds with high trapped fields. Phys. C Supercond. Its Appl. 2010, 470, 685–688. [Google Scholar] [CrossRef]
  9. Cloots, R.; Koutzarova, T.; Mathieu, J.P.; Ausloos, M. From RE-211 to RE-123. How to control the final microstructure of superconducting single-domains. Supercond. Sci. Technol. 2005, 18, R9–R23. [Google Scholar] [CrossRef]
  10. Sudhaker Reddy, E.; Rajasekharan, T. Fabrication of textured REBa2Cu3O7/RE2BaCuO5 (RE = Y, Gd) composites by infiltration and growth of RE2BaCuO5 preforms by liquid phases. Supercond. Sci. Technol. 1998, 11, 523–534. [Google Scholar] [CrossRef]
  11. Devendra Kumar, N.; Rajasekharan, T.; Gundakaram, R.C.; Seshubai, V. Extensive Nanotwinning: Origin of High Current Density to High Fields in Preform-Optimized Infiltration-Growth-Processed YBa2Cu3O7-δ Superconductor. IEEE Trans. Appl. Supercond. 2011, 21, 3612–3620. [Google Scholar] [CrossRef]
  12. Meslin, S.; Iida, K.; Babu, N.H.; Cardwell, D.A.; Noudem, J.G. The effect of Y-211 precursor particle size on the microstructure and properties of Y–Ba–Cu–O bulk superconductors fabricated by seeded infiltration and growth. Supercond. Sci. Technol. 2006, 19, 711–718. [Google Scholar] [CrossRef]
  13. Pavan Kumar Naik, S.; Pęczkowski, P.; Ogino, H.; Muralidhar, M.; Sakai, N.; Oka, T.; Nishio, T.; Murakami, M. Top-seeded infiltration growth processing of single grain (Gd, Dy)BaCuO superconductors: Nano Nb2O5 doping, enhancement of trapped field and superconducting performance. Mater. Chem. Phys. 2021, 272, 124954. [Google Scholar] [CrossRef]
  14. Kutuk, S.; Bolat, S. Analysis of structural, electrical, and levitation properties of (RE: Nd, Sm, Gd) BCO-358 superconducting ceramics prepared by a modified MPMG technique. J. Mater. Sci. 2022, 57, 1198–1214. [Google Scholar] [CrossRef]
  15. Kutuk, S.; Bolat, S. Levitation Force of (RE)BCO-358 Bulk Superconductors. AIP Conf. Proc. 2018, 2042, 020033. [Google Scholar]
  16. Yang, P.; Fagnard, J.-F.; Vanderbemden, P.; Yang, W. Magnetic shielding of a short thick GdBCO tube fabricated by the buffer aided top-seeded infiltration and growth method. Supercond. Sci. Technol. 2019, 32, 1. [Google Scholar] [CrossRef]
  17. Zhu, Y.H.; Zmorayová, K.; He, J.Y.; Zhang, Y.X.; Diko, P.; Zhou, D.F.; Yao, X. An in situ self-assembly strategy for exact-(110)-plane-controlled crystallization of high-performance YBa2Cu3O7−δ single. Ceram. Int. 2022, 48, 22196–22205. [Google Scholar] [CrossRef]
  18. Zhu, Y.H.; Xu, H.H.; Yang, J.L.; Gao, Q.; Gu, X.F.; Wang, X.C.; Zhou, D.F.; Yao, X. Sustainable recycling for epitaxial growth of YBa2Cu3O7−δ cryomagnets: From discarded buffer-crystals to superior seeds. Ceram. Int. 2023, in press. [Google Scholar] [CrossRef]
  19. Namburi, D.K.; Shi, Y.; Palmer, K.G.; Dennis, A.R.; Durrell, J.H.; Cardwell, D.A. Control of Y-211 content in bulk YBCO superconductors fabricated by a buffer-aided, top seeded infiltration and growth melt process. Supercond. Sci. Technol. 2016, 29, 034007. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Mu, Y.; Zeng, L.; Wang, M.; Yao, X. Progress in Buffer-Supported Seeding Architectures for Reliable Epitaxial Growth of REBa2Cu3O7−δ Bulk Cryomagnets with Superior Properties. Cryst. Growth Des. 2020, 20, 7533–7549. [Google Scholar] [CrossRef]
  21. Devendra Kumar, N.; Shi, Y.; Zhai, W.; Dennis, A.R.; Durrell, J.H.; Cardwell, D.A. Buffer Pellets for High-Yield, Top-Seeded Melt Growth of Large Grain Y–Ba–Cu–O Superconductors. Cryst. Growth Des. 2015, 15, 1472–1480. [Google Scholar] [CrossRef]
  22. Cardwell, D.A.; Shi, Y.; Numburi, D.K. Reliable single grain growth of (RE)BCO bulk superconductors with enhanced superconducting properties. Supercond. Sci. Technol. 2020, 33, 024004. [Google Scholar] [CrossRef]
  23. Almalki, A.; Namburi, D.K.; Ba-Abbad, M.; Dennis, A.R.; Huang, K.Y.; Almutairi, A.-A.; Durrell, J.H.; Cardwell, D.A. Processing and Properties of Bar-Shaped Single-Seeded and Multi-Seeded YBCO Bulk Superconductors by a Top-Seeded Melt Growth Technique. J. Supercond. Nov. Magn. 2016, 30, 1397–1403. [Google Scholar] [CrossRef]
  24. Goodfellow, A.; Shi, Y.-H.; Durrell, J.H.; Dennis, A.R.; Cardwell, D.A.; Grovenor, C.R.M.; Speller, S.C. Microstructural evolution in multiseeded YBCO bulk samples grown by the TSMG process. Supercond. Sci. Technol. 2016, 29, 115005. [Google Scholar] [CrossRef]
  25. Yang, W.; Chen, C.; Ma, J. Theoretical and Experimental Identification of Seeds Number on the Crystal Morphology and Physical Properties of GdBCO Bulks by Gd+011 TSIG Process. IEEE Trans. Appl. Supercond. 2019, 29, 6800904. [Google Scholar] [CrossRef]
  26. Choi, J.; Park, S.; Jun, B.; Han, Y.; Jeong, N.; Kim, B.; Sohn, J.; Kim, C. Levitation force and trapped magnetic field of multi-grain YBCO bulk superconductors. Phys. C Supercond. 2008, 468, 1473–1476. [Google Scholar] [CrossRef]
  27. Cheng, L.; Guo, L.S.; Wu, Y.S.; Yao, X.; Cardwell, D.A. Multi-seeded growth of melt processed Gd–Ba–Cu–O bulk superconductors using different arrangements of thin film seeds. J. Cryst. Growth 2013, 366, 1–7. [Google Scholar] [CrossRef]
  28. Shi, Y.; Durrell, J.H.; Dennis, A.R.; Zhang, Z.; Zhai, W.; Hari Babu, N.; Cardwell, D.A. A Comparison of 0°-0° and 45°-45° bridge-Seeded, YBCO single grains. J. Am. Ceram. Soc. 2013, 96, 1757–1762. [Google Scholar] [CrossRef]
  29. Shi, Y.H.; Durrell, J.H.; Dennis, A.R.; Cardwell, D.A. Bulk YBCO seeded with 45°–45° bridge-seeds of different lengths. Supercond. Sci. Technol. 2013, 26, 015012. [Google Scholar] [CrossRef]
  30. Li, T.Y.; Wang, C.L.; Sun, L.J.; Yan, S.B.; Cheng, L.; Yao, X.; Xiong, J.; Tao, B.W.; Feng, J.Q.; Xu, X.Y.; et al. Multiseeded melt growth of bulk Y–Ba–Cu–O using thin film seeds. J. Appl. Phys. 2010, 108, 8979. [Google Scholar] [CrossRef]
  31. Zhang, J.; Zhang, Y.; Lou, Z.; Zhang, P.; Li, C.; Yuan, J.; Zhang, X.; Peng, L.; Ma, Y.; Xu, J.; et al. Improved Flux Pinning Performance of GdBCO Superconductor Bulks by Gd3ZrO7 Additions. Acta Phys. Pol. A 2021, 140, 40–49. [Google Scholar] [CrossRef]
  32. Zhou, D.; Hara, S.; Li, B.; Xu, K.; Noudem, J.; Izumi, M. Significant improvement of trapped flux in bulk Gd–Ba–Cu–O grains fabricated by a modified top-seeded melt growth process. Supercond. Sci. Technol. 2013, 26, 015003. [Google Scholar] [CrossRef]
  33. Shi, Y.; Dennis, A.R.; Zhou, D.; Namburi, D.K.; Huang, K.; Durrell, J.H.; Cardwell, D.A. Factors Affecting the Growth of Multiseeded Superconducting Single Grains. Cryst. Growth Des. 2016, 16, 5110–5117. [Google Scholar] [CrossRef]
  34. Delamare, M.P.; Bringmann, B.; Jooss, C.; Walter, H. Influence of the seed distance on the microstructure and the superconducting properties of grain boundaries in a multi-seeded melt growth monolith. Supercond. Sci. Technol. 2002, 15, 304. [Google Scholar] [CrossRef]
  35. Ozturk, K.; Guner, S.; Abdioglu, M.; Demirci, M.; Celik, S.; Cansiz, A. An analysis on the relation between the seed distance and vertical levitation force for the multi–seeded YBCO using the modified advanced frozen image (MAFI) and experimental methods. J. Alloys Compd. 2019, 805, 1208–1216. [Google Scholar] [CrossRef]
  36. Savaskan, B.; Guner, S.B.; Yamamoto, A.; Ozturk, K. Trapped magnetic field and levitation force properties of multi-seeded YBCO superconductors with different seed distance. J. Alloys Compd. 2020, 829, 154400. [Google Scholar] [CrossRef]
  37. Kütük, S.; Bolat, S.; Öztürk, K.; Akbulut, S.; Yanmaz, E. Microstructure Study at Different Thicknesses of Disc-Shaped YBa2Cu3O7−x Prepared by FQMG Method. J. Supercond. Nov. Magn. 2011, 24, 471–474. [Google Scholar] [CrossRef]
  38. Kütük, S.; Bolat, S.; Öztürk, K.; Akbulut, S.; Çakır, O. Levitation Force Density on Different Thicknesses of YBa2Cu3O7−x Bulk Superconductor Fabricated by FQMG Process. J. Supercond. Nov. Magn. 2011, 24, 475–479. [Google Scholar] [CrossRef]
  39. Congreve, J.V.; Shi, Y.; Dennis, A.R.; Durrell, J.H.; Cardwell, D.A. Improvements in the processing of large grain bulk Y–Ba–Cu–O superconductors via the use of additional liquid phase. Supercond. Sci. Technol. 2017, 30, 015017. [Google Scholar] [CrossRef]
  40. Shi, Y.; Durrell, J.H.; Dennis, A.R.; Huang, K.; Namburi, D.K.; Zhou, D.; Cardwell, D.A. Multiple seeding for the growth of bulk GdBCO–Ag superconductors with single grain behavior. Supercond. Sci. Technol. 2017, 30, 015003. [Google Scholar] [CrossRef]
  41. Zhou, D.; Hara, S.; Li, B.; Noudem, J.; Izumi, M. Flux pinning properties of Gd–Ba–Cu–O trapped field magnets grown by a modified top-seeded melt growth. Supercond. Sci. Technol. 2014, 27, 044015. [Google Scholar] [CrossRef]
  42. Hajdová, P.; Volochová, D.; Rajňák, M.; Antal, V.; Diko, P. Growth, Microstructure, and Properties of GdBCO–Ag Superconductor. IEEE Trans. Appl. Supercond. 2019, 29, 6800205. [Google Scholar] [CrossRef]
  43. Zhou, D.; Xu, K.; Hara, S.; Li, B.; Deng, Z.; Tsuzuki, K.; Izumi, M. MgO buffer-layer-induced texture growth of RE–Ba–Cu–O bulk. Supercond. Sci. Technol. 2012, 25, 025022. [Google Scholar] [CrossRef]
  44. Endo, A.; Chauhan, H.S.; Egi, T.; Shiohara, Y. Macrosegregation of Y2Ba1Cu1O5 particles in Y1Ba2Cu3O7−δ crystals grown by an undercooling method. J. Mater. Res. 1996, 11, 795–803. [Google Scholar] [CrossRef]
  45. Shi, Y.; Gough, M.; Dennis, A.R.; Durrell, J.H.; Cardwell, D.A. Distribution of the superconducting critical current density within a Gd–Ba–Cu–O single grain. Supercond. Sci. Technol. 2020, 33, 1. [Google Scholar] [CrossRef]
Figure 1. Schematic assembly of pellets.
Figure 1. Schematic assembly of pellets.
Micromachines 14 00987 g001
Figure 2. The temperature curve of the superconductor bulk SA and SB.
Figure 2. The temperature curve of the superconductor bulk SA and SB.
Micromachines 14 00987 g002
Figure 3. Macromorphology of SA (a) and SB (b). The blue dotted line indicates the growth facet line on the top surface of the sample, and the red arrow indicates a slight drift between the two seed crystals. (c) Schematic diagram of the positions of two specimens cut from the GdBCO superconductor bulk.
Figure 3. Macromorphology of SA (a) and SB (b). The blue dotted line indicates the growth facet line on the top surface of the sample, and the red arrow indicates a slight drift between the two seed crystals. (c) Schematic diagram of the positions of two specimens cut from the GdBCO superconductor bulk.
Micromachines 14 00987 g003aMicromachines 14 00987 g003b
Figure 4. The top surface trapped field (3D map and 2D map) of SA and SB measured in a liquid nitrogen atmosphere, (a) 2D map of SA; (b) 2D map of SB; (c) 3D map of SA; (d) 3D map of SB, respectively.
Figure 4. The top surface trapped field (3D map and 2D map) of SA and SB measured in a liquid nitrogen atmosphere, (a) 2D map of SA; (b) 2D map of SB; (c) 3D map of SA; (d) 3D map of SB, respectively.
Micromachines 14 00987 g004aMicromachines 14 00987 g004b
Figure 5. Bottom surface trapped field (3D map and 2D map) of SA and SB measured in liquid nitrogen atmosphere, (a) 2D map of SA; (b) 2D map of SB; (c) 3D map of SA; (d) 3D map of SB, respectively.
Figure 5. Bottom surface trapped field (3D map and 2D map) of SA and SB measured in liquid nitrogen atmosphere, (a) 2D map of SA; (b) 2D map of SB; (c) 3D map of SA; (d) 3D map of SB, respectively.
Micromachines 14 00987 g005
Figure 6. Temperature dependence of magnetization of specimens cut below the seed crystals of SA (a) and SB (b), respectively.
Figure 6. Temperature dependence of magnetization of specimens cut below the seed crystals of SA (a) and SB (b), respectively.
Micromachines 14 00987 g006
Figure 7. The critical current density curve measured by cutting the specimens below the seed crystal of SA (a) and SB (b), respectively.
Figure 7. The critical current density curve measured by cutting the specimens below the seed crystal of SA (a) and SB (b), respectively.
Micromachines 14 00987 g007
Figure 8. Plots of the JC in self-field of specimens from different positions under seed crystal.
Figure 8. Plots of the JC in self-field of specimens from different positions under seed crystal.
Micromachines 14 00987 g008
Figure 9. Scanning electron microscope photos of 500 times magnification of cut specimens of SA (100/100) and SB (110/110) at positions b1, b3 and b5.
Figure 9. Scanning electron microscope photos of 500 times magnification of cut specimens of SA (100/100) and SB (110/110) at positions b1, b3 and b5.
Micromachines 14 00987 g009
Figure 10. Scanning electron microscope photos of 5000 times magnification of cut specimens of SA (100/100) and SB (110/110) at positions b1, b3 and b5.
Figure 10. Scanning electron microscope photos of 5000 times magnification of cut specimens of SA (100/100) and SB (110/110) at positions b1, b3 and b5.
Micromachines 14 00987 g010
Figure 11. The size distribution of the Gd211 particles in the GdBCO superconductor bulks at positions b1, b3 and b5. The red solid curves represent the fitted curves obtained by fitting the size distribution of the Gd211 particles. The average size values of the Gd211 particles (AVG) obtained from the peak position of each fitted curve are also given, (a) SA-b1; (b) SA-b3; (c) SA-b5; (d) SB-b1; (e) SB-b3; (f) SB-b5, respectively.
Figure 11. The size distribution of the Gd211 particles in the GdBCO superconductor bulks at positions b1, b3 and b5. The red solid curves represent the fitted curves obtained by fitting the size distribution of the Gd211 particles. The average size values of the Gd211 particles (AVG) obtained from the peak position of each fitted curve are also given, (a) SA-b1; (b) SA-b3; (c) SA-b5; (d) SB-b1; (e) SB-b3; (f) SB-b5, respectively.
Micromachines 14 00987 g011
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

Zhang, Y.; Li, C.; Lou, Z.; Zhang, P.; Zhang, Y.; Shen, S.; Ruan, G.; Zhang, J. The Performance of the Two-Seeded GdBCO Superconductor Bulk with the Buffer by the Modified TSMG Method. Micromachines 2023, 14, 987. https://doi.org/10.3390/mi14050987

AMA Style

Zhang Y, Li C, Lou Z, Zhang P, Zhang Y, Shen S, Ruan G, Zhang J. The Performance of the Two-Seeded GdBCO Superconductor Bulk with the Buffer by the Modified TSMG Method. Micromachines. 2023; 14(5):987. https://doi.org/10.3390/mi14050987

Chicago/Turabian Style

Zhang, Yufeng, Chunyan Li, Ziwei Lou, Penghe Zhang, Yan Zhang, Shuangyuan Shen, Guanjie Ruan, and Jiaying Zhang. 2023. "The Performance of the Two-Seeded GdBCO Superconductor Bulk with the Buffer by the Modified TSMG Method" Micromachines 14, no. 5: 987. https://doi.org/10.3390/mi14050987

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