Next Article in Journal
Copper Nitride: A Versatile Semiconductor with Great Potential for Next-Generation Photovoltaics
Previous Article in Journal
Design and Preparation of Anti-Reflection Nanoarray Structure on the Surface of Space Solar Cell Glass Cover
Previous Article in Special Issue
Joining Superconducting MgB2 Parts by Spark Plasma Sintering: A New Technique with High Potential for Manufacturing Future Superconducting Devices
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Orthorhombic YBa2Cu3O7−δ Superconductor with TiO2 Nanoparticle Addition: Crystal Structure, Electric Resistivity, and AC Susceptibility

by
Fatma Barood
1,
Mohd Mustafa Awang Kechik
1,*,
Tan Sin Tee
1,
Chen Soo Kien
1,
Lim Kean Pah
1,
Kai Jeat Hong
1,
Abdul Halim Shaari
1,
Hussein Baqiah
2,
Muhammad Khalis Abdul Karim
1,
Muhammad Kashfi Shabdin
1,
Khairul Khaizi Mohd Shariff
3,
Azhan Hashim
4,
Nurbaisyatul Ermiza Suhaimi
4 and
Muralidhar Miryala
5
1
Laboratory of Superconductor & Thin Films, Department of Physics, Faculty of Science, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
2
Shandong Key Laboratory of Biophysics, Institute of Biophysics, Dezhou University, No. 566 University Rd. West, Dezhou 253023, China
3
Microwave Research Institute, Universiti Teknologi MARA, Shah Alam 40450, Selangor, Malaysia
4
Faculty of Applied Sciences, Universiti Teknologi MARA Pahang, Bandar Pusat Jengka 26400, Pahang, Malaysia
5
Materials for Energy and Environmental Laboratory, Superconducting Materials, Shibaura Institute of Technology, 3 Chome-7-5 Toyosu, Koto, Tokyo 135-8548, Japan
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(6), 1093; https://doi.org/10.3390/coatings13061093
Submission received: 3 May 2023 / Revised: 6 June 2023 / Accepted: 9 June 2023 / Published: 13 June 2023
(This article belongs to the Special Issue New Advance in Superconductor and Superconducting Thin Films)

Abstract

:
This article reports the effect of a nanoscale addition of TiO2 on the structure and superconducting parameters of the high-temperature superconductor YBa2Cu3O7-δ (Y123). Polycrystalline compounds of Y123 with different percentages of TiO2, x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0, were fabricated using the thermal treatment method. An analysis using X-ray diffraction confirmed the formation of Y123 phases for all composite samples. Field-emission scanning electron microscopy (FESEM) analysis revealed the growth of grain size and decrease in porosity, with a sign of partial melting of grains for the samples with TiO2 addition. The magnetic and electric transport properties were investigated using AC susceptibility measurement and the four-probe method, respectively. It was observed that the superconducting transition temperature, Tc-onset, for a pure sample determined by ACS and 4PP was 95.6 K and 95.4 K, respectively. These values were found to decrease with the addition of TiO2, while the superconducting transition (∆Tc) improved with TiO2 addition except for the sample at x = 0.2 wt.%, which showed the broadest transition width. The sharpest superconducting transition (∆Tc) was observed for the sample at x = 1.0 wt.%, indicating that the addition of TiO2 nanoparticles is expected to serve as artificial pinning centres and strengthen the connection among the grains in the Y123 ceramic.

1. Introduction

Since Chu [1] discovered the YBa2Cu3O7−δ (YBCO) superconductor in 1987, there has been considerable interest in the physics and applications of this system, leading to many research efforts. This compound is a type-II superconductor and exhibits a critical temperature transition, Tc, that can reach 92 K, which is higher than liquid nitrogen’s temperature (Tc = 77 K), making it attractive for the demands of practical applications [2,3]. The requirements of the Y123 applications would not be achieved just by a high Tc; a larger critical current density, Jc, in the presence of a magnetic field is crucial for the Y123 compound’s optimum performance. However, poor vortex pinning and weak grain couplings contribute to low Jc, hence limiting the use of Y123 material [4,5]. By introducing effective flux-line pinning, it is possible to prevent vortex motion and enhance the critical current density. In addition, it will be essential to fabricate samples with artificial pinning sites in order to improve the critical current density.
It has been widely reported that the incorporation of nano-sized precursors into YBCO bulk could be a significant way to enhance Jc with strong pinning, mainly at high magnetic fields. The flux pinning enhancement can be optimised when the inclusion of nano-sized particles is in the range of the superconducting coherence length of YBCO (2–4 nm) [6]. A significant growth in terms of improvement of the interconnectivity and the flux pinning has been reported by adding various impurities to YBCO, e.g., Y2O3, DY2O3, CO, WO3, and AL2O3 [7,8,9,10,11]. These metal oxides can act as artificial pinning centres for high critical current density. Among various nanostructure materials, titanium oxide nanoparticles (TiO2) have attracted the interest of several researchers studying superconductivity. Recently, TiO2 has demonstrated its potential as a highly effective candidate for reinforcing superconductor ceramics. Therefore, several studies were conducted using TiO2 as an additive to some of the superconducting compounds, such as MgB2 [12,13,14], BPSCCO [15], Bi-2212, and Bi-2223 [16]. The effect of TiO2 addition on the YBCO bulk has been reported a few times using a variety of synthesis methods. It has been found that the characteristics of the YBCO material are significantly influenced by the synthesis method and the amount of TiO2 nanoparticles. Based on the various addition levels, significant results have been achieved by many research groups with regards to both critical temperature and critical current. Ghahramani et al. [17] studied the effect of TiO2 on the superconducting parameters of Y123 prepared via the solid-state method. The Tc increased from 95.50 K for a pure sample to 97.77 K at x = 0.30 wt.% of TiO2, whereas Jc decreased with the TiO2 addition compared to the pure sample. Rejith et al. [18] reported the improvement in Jc of Y123 by TiO2 inclusion, while Tc was slightly decreased from 92 to 90.5 K. It was also found that the flux pinning force was eight times stronger than that in the pure YBCO sample. Hannachi et al. [19] investigated the impact of TiO2 nanoparticle addition on the intra-granular and inter-granular properties of Y123. The addition of TiO2 served as efficient pinning centres in the YBCO system, which consequently improved the connectivity between the grains of YBCO. The calculated critical current density, J (0)c, inter, was enhanced and was higher by three at x = 0.1 wt.% of the TiO2 addition sample compared to the non-added one. Kebbede et al. [20] studied the anisotropic grain growth of Y123 with (2.5 and 5 mol%) TiO2 nanoparticles incorporated via the sol-gel coating method. The critical onset temperature, Tc-onset, was decreased for the 2.5 mol% TiO2-coated sample, whereas the normal-state resistivity was higher with a 5 mol% addition due to the presence of the Y211 and BaCuO2 secondary phases. Therefore, the anisotropic grain growth of YBCO was enhanced at lower levels of TiO2 addition. The use of TiO2 nanoparticles as an additive in YBCO has proven to be an effective way of enhancing its performance. Therefore, more research is required to accurately validate the effect of TiO2 addition.
Investigations into this material are still being conducted, in particular based on various doping rates and novel technological uses. Generally, the method of preparation and the nature of the addition have a significant impact on the properties of superconducting materials. The main focus of this study is to apply the thermal treatment technique to prepare Y123 doped with TiO2 nanoparticles. To our knowledge, however, no literature has reported the use of TiO2 as an addition to YBCO by using thermal treatment as a preparation method. This method was previously found to have the advantages of being simple, inexpensive, and capable of producing fine powders [21,22,23,24]. Thus, the objective of the present study is to investigate the influence of various amounts of TiO2 nanoparticles (x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%) on the structure and superconductivity of the Y123 superconductor prepared using the thermal treatment method. Investigations were conducted on phase formation, chemical composition, AC susceptibility, and electrical resistivity.

2. Materials and Methods

The samples of Y123 pellets with different values of TiO2 were synthesised using a thermal treatment method that uses PVP as a capping agent to reduce agglomeration [25]. The starting compositions of high-purity raw materials from Haverhill, MA, Alfa Aesar, Y(NO3)2.6H2O (99.99%), Ba (NO3)2 (99.95%), and Cu (NO3)2.2.5H2O (98%) powders with the appropriate stoichiometric atomic ratio of 1:2:3, were dissolved in 300 mL of deionised water and 6 g of PVP (C6 H9NO)6. The mixed solution was heated at 80 °C for 2 h with the help of a magnetic stirrer hot plate. The solution was dried in an oven at 110 °C for 24 h to allow for water evaporation. The resultant green gel was ground for around 1 h using a mortar and pestle until a fine powder formed. Then, the powder was pre-calcined at 600 °C for 4 h using a box furnace and reground again for 15 min before a second calcination was conducted in a double-tube furnace of Model CMTS TF 40/360 at 910 °C for 24 h. The obtained powder after second calcination was mixed with the required wt.% of TiO2 during the grinding process to be pressed into 13 mm diameter pellets and sintered at 980 °C for 24 h. Extra annealing was conducted at 650 °C for 12 h in an oxygen-rich environment, followed by cooling at 1 °C per hour to prevent oxygen deprivation.
The sintered samples were examined using X-ray diffraction (XRD, Xpert Pro Panalytical Philips DY 1861 diffractometer, Phillips, Eindhoven, The Netherlands) with a CuKα source from 2θ = 20° to 80° to evaluate phase identification and the crystal structure of the sample. The microstructure analysis was performed using a field-emission scanning electron microscope along with an energy-dispersive X-ray spectrometer for quantitative analyses (FESEM, FEI Nova NanoSEM 230, Thermo Fisher Scientific, Waltham, MA, USA). Measurement of electric transport properties was conducted in the temperature range 30–280 K with the four-point probe technique using a digital nanovoltmeter (Keithley, Model 2182A, Cleveland, OH, USA) and a DC precision power source (Keithley, Model 6221, Cleveland, OH, USA). The measurements involved determine the values of onset critical temperature, Tc-onset, and offset critical temperature, Tc-offset. The temperature dependence of AC susceptibility was investigated using a susceptometer from Cryo Industry, model number REF-1808-ACS (ACS, CryonBIND T, CryoBIND, Zagred, Croatia). The frequency of the AC signal was 295 Hz, and a magnetic field of 5 Oe was applied.

3. Result and Discussion

3.1. XRD Analysis

Figure 1a describes the Rietveld refinement and fitted XRD profile of Y123/(TiO2)x patterns, where x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%. The XRD diffraction peaks for all samples exhibited an orthorhombic perovskite structure and belong to space group Pmmm 47 symmetry with ICSD No. 98-002-1157. The (013 and 103) diffraction peaks, which are characteristic of the YBCO phase, can be observed clearly in both non-added and TiO2-added samples. A small amount of impurity phase (indicated as Y211 in the figure) was observed for all added TiO2 patterns. However, there were no detected peaks attributed to TiO2 addition even with a high percentage of TiO2 nanoparticles which accords with previous outcomes stated by [18,26]. The absence of TiO2 nanoparticle structure in the XRD pattern can be attributed to the mixing effect behaviour, which is due to the relatively higher concentration of the Y123 phase compared to the TiO2. The strong diffraction peaks from the Y123 components can overshadow or mask the weaker diffraction peaks from TiO2 nanoparticles, making it challenging to observe the TiO2 crystal structure in the XRD pattern. Our results showed that the diffraction peaks increased in intensity for all TiO2-added samples, particularly at x = 0.4 wt.% TiO2 concentration, with a noticeable shift of the peak position to the high angle, as presented in Figure 1b. This might be related to thermal cycle stress or the powder pelletising process [27].
The variation of the lattice parameters a, b, and c with TiO2 concentration is shown in Table 1. It is noticed that the lattice parameter a basically remained the same, while the values of the lattice parameters b and c increased with TiO2 addition level. These differences in the lattice constants a and b resulted in increased orthorhombicity (ab)/ (a + b) of the superconducting phase. The crystallite size of all added samples was reduced as a result of adding TiO2, which limited the crystallite growth of the Y123 ceramic. The oxygen stoichiometry, which is known to play a crucial role in the superconducting properties of YBCO [28], was calculated for all sintered samples. The oxygen content of Y123 samples was estimated from the c-axis value using the relation 7−δ = 75.25 − 5.856 c [29]. It was noticed that the oxygen content decreased for all Y123/TiO2 composites. The relation between parameter c-axis and oxygen level is also consistent with the typical behaviour of c-axis in the orthorhombic phase Y123, as it is well known that the c-axis parameter increases when the oxygen content decreases [29]. The crystallite size, orthorhombicity, and oxygen content values are tabulated in Table 1.

3.2. Microstructure Analysis

The FESEM analysis was conducted to further investigate the impact of TiO2 nanoparticle addition on the microstructure of YBCO samples. Figure 2a–f shows the surface morphology of the Y123/(TiO2)x composite along with EDX spectra and histograms of the average grain size distribution. As can be seen, the surface morphology was modified under TiO2 inclusion compared to pure one. The grain boundaries were defined as a granular microstructure with shaped rectangular grains distributed uniformly among the sample. The grains were closely packed and elongated with the addition of TiO2. In addition, a sign of partial melting was observed in the grains, leading to the presence of a poorly defined grain boundary. This indicated the fusion of several neighbouring grains, leading to an increase in grain size during melting [30]. The average grain size was 1.59, 2.82, 2.50, 2.74, 3.04, and 2.48 μm for the samples with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0, respectively. It is clearly evident that the addition of TiO2 increased the grain size of different sintered Y123 samples. It is thought that in ceramic materials, higher sintering temperatures cause partial melting and merging of the grain boundaries, which promote grain size growth [31,32]. As a result, no pores were detected in these composite samples because the grains of Y123 continued to grow and filled the gaps among the grains. On the other hand, the increase in grain size with TiO2 can create better inter-grain connections, allowing the flow of inter-granular electric current. Hence, the transport critical current Jc is expected to increase with TiO2 additive.
Figure 2a–f displays EDX spectra along with a histogram of the average grain size distribution for each sample of Y123/(TiO2)x composites at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%. The EDX analysis confirmed the presence of the elements Y, Ba, Cu, and O for all the samples, as reported in Table 2. The atomic ratios for Y:Ba: Cu were matched to the main composition of 1:2:3. The spectra graph of EDX also detected Ti ion peaks for 0.6 and 1.0 wt.%, while it was not detected in the other percentages. The limited resolution in space of the analysis and the limited sample area selected for EDX analysis may result in variations in the detected elemental composition, as the analysed regions are only point-sized. Consequently, there is a possibility of missing Ti particles or areas that contain Ti but may be present in other regions of the surface since the sampling spots were selected in a random manner. It is also common to observe in bulk samples that the grinding of all components during the addition process is thought to result in their random separation. These EDX results show that the Y123 phase dominates in all prepared samples, which is in accord with XRD investigations.

3.3. Alternating Current Susceptibility (ACS) Measurement

Figure 3 displays the AC susceptibility of real, χ′, and imaginary, χ″, parts for the Y123 superconductor with different percentages of TiO2 at x = (0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%), measured at Hac = 5 Oe and f = 295 Hz. The values of the phase lock-in temperature, Tcj, and the diamagnetic onset temperature, Tc-onset, were determined from the plot of the peaks dχ´/dT versus T shown in Figure 4. It is noticed that for all added TiO2 samples, the Tc-onset showed a slight decrease with TiO2 inclusion, while Tcj was greatly increased from 80.1 k for the pure Y123 samples to 91.5 K for 1.0 wt.% TiO2. The increase in Tcj implies that the coupling between the grains was further enhanced with TiO2. Based on the real part (χ′-T), it is observed that the transition became sharper with TiO2 addition compared to the non-added sample, which showed tow-step transition. This indicates a strong granular nature of these samples as the addition of TiO2 nanoparticles increased. Furthermore, having a sharp transition in HTSC materials is evidence that the critical current density would be enhanced [33].
The Imaginary part, χ″ in the (χ″-T) susceptibility displays the AC losses. Unlike the pure sample, two peaks were observed for TiO2 samples, denoted as Tpm and Tpg, which correspond to inter-granular and intra-granular, respectively. The values of Tpm and Tpg are listed in Table 3. The loss peaks, Tpg, were found to occur at a higher temperature in samples with TiO2 addition, revealing that dissipation for samples containing TiO2 begins at higher temperatures. The sample with x = 0.6 exhibited the highest Tpm, whereas for other TiO2 contents, Tpm slightly decreased as TiO2 increased. This can be attributed to the inhibition of inter-grain coupling within the superconducting grains. It is known that the AC magnetic characteristics of high-Tc superconductors are caused by shielding currents that flow either inside the grains (intra-granular currents) or between grain boundaries (inter-granular currents) [34]. It should be noted that all Y123/TiO2 composite samples showed weak peaks and broadening of the inter-granular peaks when compared to the pure sample. This indicated field penetration (due to granular quality) as well as hysteretic losses between grains [35]. The values of the inter-granular current density Jc (Tpm) for all samples were calculated using the Bean model equation Jc (Tpm) = H/(ab)0.5 [36]. The value of Jc (Tpm) for all sintered samples was around 19 to 22 A/cm2.
Due to the granular nature of ceramic superconductors, YBCO can be modelled as an array of weak Josephson junctions, with grain coupling occurring through Josephson currents. The maximum inter-grain Josephson current, I0, can be determined from the obtained values of Tcj and Tc-onset using the following equations assumed by the Ambegaokar–Baratoff theory [37,38].
I 0 = 1.57 × 10 - 8 T c - o n s e t 2 T c - o n s e t - T c j
E j = h 4 π e   I 0
where h is Planck’s constant, e is the electron’s charge, and I0 is the maximum Josephson current.
As revealed in Table 3, there was a remarkable increase in the values of I0 and Ej for all samples with TiO2 inclusion, with the highest value for the sample containing 0.4 wt.% TiO2 addition. This result indicates that TiO2 incorporation in YBCO bulk enhances the inter-grain coupling of the grains. Table 3 summarises the data for Tc-onset, Tp, Tcj, Jc (Tpm), I0, and Ej.

3.4. Electrical Resistivity Measurement

Normalised resistivity as a function of temperature is displayed in Figure 5a for x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.% TiO2 added to Y123 bulk. The resistivity measurements were performed on bulk Y123 using the standard four-point probe technique in a temperature range of 30 to 280 K. All samples displayed a roughly linear drop in resistivity when cooled from room temperature to the critical onset temperature of the superconducting transition. Thus, a metal behaviour was confirmed for all Y123/TiO2 compounds. Analysing these curves showed a gradual degradation of the critical onset temperature, Tc-onset, up to 0.8 wt.% TiO2 samples and then an increase for the sample at x = 1.0 wt.% TiO2. This decrease in Tc-onset, in comparison to pure Y123, could be due to secondary phase formation, the solidification process, or the effect of microscopic inhomogeneity in the Y123/TiO2 substance. The reduction in oxygen content in the CuO chains may be another satisfactory reason related to Tc-onset degradation with TiO2 addition [11,39]. The onset of superconducting transition temperature, Tc-onset, for a pure sample was observed at 95.6 K, which is consistent with the Tc-onset value obtained by AC susceptibility (95.4 K). Notably, the decline in Tc-onset measured by AC susceptibility and the four-point probe method has an almost similar trend, as shown in Figure 6, which displays the comparison results of Tc-onset measured using the two methods. The electrical resistivity started to drop at 95.6, 93.1, 92.6, 92.5, 91.1, and 94.7 K, whereas it vanished at 91.7, 84.6, 89.1, 89, 87.2, and 91.7 K for x = (0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%) TiO2, respectively. The residual resistivity, ρ0, was decreased for all added samples, indicating improvement in weak links among the grains with TiO2 nanoparticle inclusion. The sample at x = 1.0 wt.% was found to have the lowest value of ρ0. All determined values of the residual resistivity are presented in Table 4.
More information about the transition can be obtained from the dρ/dT curves shown in Figure 5b. The transition width, ΔTc, was around 3–4 K except at x = 0.2 wt.%, which exhibited a broadest ΔTc of 8 K. Notably, the dρ/dT peaks of the high-level concentration of TiO2 (0.6–1.0 wt.%) exhibited high and sharp peaks. It is believed that the improvement in the superconducting transition width is because of the modification effect of the TiO2 inclusion on the grain boundary, which resulted in grain connectivity enhancement. Note that the sample with the highest concentration (1.0 wt.% TiO2) showed the narrowest transition width, ΔTc (3 K), and its peak was slightly shifted to a higher temperature. This result is consistent with the ACS result, as the same sample exhibited the sharpest and highest peak. These results suggest that a positive impact can be achieved as the concentration of TiO2 increases. The hole concentration, p, in the CuO2 plane was calculated using a well-known relation between Tc and the charge density [40].
p = 0.16 - 1 - T c - o n s e t T c m a x / 82.6 0.5
where T c m a x in this study is 95.6 K for the YBCO phase, and p is the number of the hole concentration. The hole concentration decreased for all doped TiO2 samples from 0.160 to 0.138. The change in oxygen content in the Y123 system may be responsible for this reduction in the charge carrier [41,42]. The summary of Tc-onset (K), Tc-offset (K), ∆Tc (K), and hole concentration, p, is listed in Table 4.

4. Conclusions

The composite ceramic Y123/(TiO2)x (x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%) synthesised by the thermal treatment method and annealed at 650 °C in a flowing oxygen atmosphere was successfully prepared. XRD analysis showed that an orthorhombic Y123 single-phase with a crystal structure was observed for all samples, with a trace to the Y211 secondary phase for the samples with TiO2 addition. There was a slight dependence of the c-axis on TiO2 concentration, which influenced the oxygen content (7-δ). The orthorhombicity increased with TiO2 addition, indicating better oxygen ordering. The findings of the FESEM analysis indicated that the incorporation of TiO2 nanoparticles led to an increase in the average grain size, resulting in a reduction in pores and an overall increase in sample compactness. Electrical resistivity and AC susceptibility measurements showed that Tc-onset was slightly decreased for all Y123/TiO2 samples. The superconducting transition, ΔTc, was improved with TiO2 inclusion and was the sharpest at x = 1.0 wt.%. The residual resistivity, ρ0, decreased for all added samples, indicating a weak link improvement among the grains due to TiO2 nanoparticle inclusion. From the AC susceptibility measurement, several significant parameters were determined and discussed. The results showed a significant increase in the values of Io and Ej for all samples with TiO2. This suggests that the addition of TiO2 enhanced the inter-grain coupling by improving grain connectivity as it incorporated between the grains and served as efficient pinning centres inside the Y123 ceramic. These results prove the positive impact of TiO2 on Y123 performance.

Author Contributions

Conceptualization, F.B., M.M.A.K. and M.M.; methodology, F.B. and K.K.M.S.; software, F.B., N.E.S. and A.H.; validation, F.B., C.S.K. and L.K.P.; formal analysis, F.B., T.S.T. and A.H.S.; investigation, F.B., M.M.A.K. and M.M.; resources, F.B. and M.M.A.K.; data curation, F.B., K.J.H., M.K.A.K. and M.K.S.; writing—original draft preparation, F.B.; writing—review and editing, M.M.A.K., M.M. and H.B.; project administration, M.M.A.K. and M.M.; funding acquisition, M.M.A.K. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Higher Education (MOHE) Malaysia under FRGS Grant No. FRGS/1/2017/STG02/UPM/02/4 and Universiti Putra Malaysia IPS Grant No. 9666600.

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. The data are not publicly available due to ethical restrictions.

Acknowledgments

Additionally, this work was partly supported partly supported by Japan Science and Technology (JST) for the Advanced Project Based Learning (aPBL), Shibaura Institute of Technology. The authors would like to thank the valuable support and discussions by all members of the Superconductor and Thin Films Laboratory of Universiti Putra Malaysia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chu, C.W. Superconductivity above 90 K. Proc. Natl. Acad. Sci. USA 1987, 84, 4681–4682. [Google Scholar] [CrossRef] [Green Version]
  2. Hassenzahl, W.V. Superconductivity, an enabling technology for 21st century power systems? IEEE Trans. Appl. Supercond. 2001, 11, 1447–1453. [Google Scholar] [CrossRef]
  3. Kelley, N.; Nassi, M.; Masur, L. Application of HTS wire and cables to power transmission: State of the art and opportunities. In Proceedings of the IEEE Power Engineering Society Transmission and Distribution Conference, Columbus, OH, USA, 28 January 2001–1 February 2001; Volume 2, pp. 448–454. [Google Scholar] [CrossRef]
  4. Bishop, D.J.; Gammel, P.L.; Huse, D.A.; Murray, C.A. Magnetic Flux-Line Lattices and Vortices in the Copper Oxide Superconductors. Science 1992, 255, 165–172. [Google Scholar] [CrossRef]
  5. Larbalestier, D. Superconductor Flux Pinning and Grain Boundary Control. Science 1996, 274, 736–737. [Google Scholar] [CrossRef]
  6. Haugan, T.; Barnes, P.N.; Wheeler, R.; Meisenkothen, F.; Sumption, M. Addition of nanoparticle dispersions to enhance flux pinning of the YBa2Cu3O7-x superconductor. Nature 2004, 430, 867–870. [Google Scholar] [CrossRef]
  7. Çakır, B.; Aydıner, A.; Basoglu, M.; Yanmaz, E.; Aydiner, A. The Effect of Y2O3 on AC Susceptibility Measurements of MPMG YBCO Superconductor. J. Supercond. Nov. Magn. 2013, 26, 937–941. [Google Scholar] [CrossRef]
  8. Almessiere, M.A.; Slimani, Y.; Hannachi, E.; Algarni, R.; Ben Azzouz, F. Impact of Dy2O3 nanoparticles additions on the properties of porous YBCO ceramics. J. Mater. Sci. Mater. Electron. 2019, 30, 17572–17582. [Google Scholar] [CrossRef]
  9. Slimani, Y.; Almessiere, M.; Hannachi, E.; Mumtaz, M.; Manikandan, A.; Baykal, A.; Ben Azzouz, F. Improvement of flux pinning ability by tungsten oxide nanoparticles added in YBa2Cu3Oy superconductor. Ceram. Int. 2019, 45, 6828–6835. [Google Scholar] [CrossRef]
  10. Dadras, S.; Falahati, S.; Dehghani, S. Effects of graphene oxide doping on the structural and superconducting properties of YBa2Cu3O7−δ. Phys. C Supercond. Its Appl. 2018, 548, 65–67. [Google Scholar] [CrossRef]
  11. Mellekh, A.; Zouaoui, M.; Ben Azzouz, F.; Annabi, M.; Ben Salem, M. Nano-Al2O3 particle addition effects on Y Ba2Cu3Oy superconducting properties. Solid State Commun. 2006, 140, 318–323. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Zhou, S.; Wang, X.; Dou, S.X. Effect of addition of nanoparticle TiO2/SiO2 on the superconducting properties of MgB2. Phys. C Supercond. Its Appl. 2008, 468, 1383–1386. [Google Scholar] [CrossRef]
  13. Prokhorov, V.G.; Svetchnikov, V.L.; Park, J.S.; Kim, G.H.; Lee, Y.P.; Kang, J.-H.; Khokhlov, V.A.; Mikheenko, P. Flux pinning and the paramagnetic Meissner effect in MgB2 with TiO2 inclusions. Supercond. Sci. Technol. 2009, 22, 045027. [Google Scholar] [CrossRef]
  14. Kishan, H.; Awana, V.; de Oliveira, T.; Alam, S.; Saito, M.; de Lima, O. Superconductivity of nano-TiO2-added MgB2. Phys. C Supercond. Its Appl. 2007, 458, 1–5. [Google Scholar] [CrossRef]
  15. Pham, A.T.; Tran, D.T.; Vu, L.H.; Chu, N.T.; Thien, N.D.; Nam, N.H.; Binh, N.T.; Tai, L.T.; Hong, N.T.; Long, N.T.; et al. Effects of TiO2 nanoparticle addition on the flux pinning properties of the Bi1.6Pb0.4Sr2Ca2Cu3O10+δ ceramics. Ceram. Int. 2022, 48, 20996–21004. [Google Scholar] [CrossRef]
  16. Hamid, N.A.; Abd-Shukor, R. Effects of TiO2 addition on the superconducting properties of Bi-Sr-Ca-Cu-O system. J. Mater. Sci. 2000, 35, 2325–2329. [Google Scholar] [CrossRef]
  17. Ghahramani, S.; Shams, G.; Soltani, Z. Comparative Investigation of the Effect of Titanium Oxide Nanoparticles on Some Superconducting Parameters of Y3Ba5Cu8O18±δ and Y1Ba2Cu3O7−δ Composites. J. Electron. Mater. 2021, 50, 4727–4740. [Google Scholar] [CrossRef]
  18. Rejith, P.; Vidya, S.; Thomas, J. Improvement of Critical Current Density in YBa2Cu3O7-δ Superconductor with Nano TiO2 Addition. Mater. Today Proc. 2015, 2, 997–1001. [Google Scholar] [CrossRef]
  19. Hannachi, E.; Slimani, Y.; Ben Azzouz, F.; Ekicibil, A. Higher intra-granular and inter-granular performances of YBCO superconductor with TiO2 nano-sized particles addition. Ceram. Int. 2018, 44, 18836–18843. [Google Scholar] [CrossRef]
  20. Kebbede, A.; Parai, J.; Carim, A.H. Sol-gel coating of YBa2Cu3O7−x with TiO2 for enhanced anisotropic grain growth. Energy 1993, 51, 2845–2851. [Google Scholar]
  21. Yusuf, N.N.M.; Kechik, M.M.A.; Baqiah, H.; Kien, C.S.; Pah, L.K.; Shaari, A.H.; Jusoh, W.N.W.W.; Sukor, S.I.A.; Dihom, M.M.; Talib, Z.A.; et al. Structural and Superconducting Properties of Thermal Treatment-Synthesised Bulk YBa2Cu3O7−δ Superconductor: Effect of Addition of SnO2 Nanoparticles. Materials 2019, 12, 92. [Google Scholar] [CrossRef] [Green Version]
  22. Dihom, M.M.; Shaari, A.H.; Baqiah, H.; Al-Hada, N.M.; Kien, C.S.; Azis, R.S.; Kechik, M.M.A.; Talib, Z.A.; Abd-Shukor, R. Microstructure and superconducting properties of Ca substituted Y(Ba1−Ca)2Cu3O7−δ ceramics prepared by thermal treatment method. Results Phys. 2017, 7, 407–412. [Google Scholar] [CrossRef]
  23. Dihom, M.M.; Shaari, A.H.; Baqiah, H.; Al-Hada, N.M.; Talib, Z.A.; Kien, C.S.; Azis, R.S.; Kechik, M.M.A.; Pah, L.K.; Abd-Shukor, R. Structural and superconducting properties of Y(Ba1-K)2Cu3O7-δ ceramics. Ceram. Int. 2017, 43, 11339–11344. [Google Scholar] [CrossRef]
  24. Kamarudin, A.N.; Kechik, M.M.A.; Abdullah, S.N.; Baqiah, H.; Chen, S.K.; Karim, M.K.A.; Ramli, A.; Lim, K.P.; Shaari, A.H.; Miryala, M.; et al. Effect of Graphene Nanoparticles Addition on Superconductivity of YBa2Cu3O7~δ Synthesized via the Thermal Treatment Method. Coatings 2022, 12, 91. [Google Scholar] [CrossRef]
  25. Dihom, M.M.; Shaari, A.H.; Baqiah, H.; Al-Hada, N.M.; Chen, S.K.; Azis, R.S.; Kechik, M.M.A.; Abd-Shukor, R. Effects of Calcination Temperature on Microstructure and Superconducting Properties of Y123 Ceramic Prepared Using Thermal Treatment Method. Solid State Phenom. 2017, 268, 325–329. [Google Scholar] [CrossRef]
  26. Slimani, Y.; Hannachi, E.; Ekicibil, A.; Almessiere, M.; Ben Azzouz, F. Investigation of the impact of nano-sized wires and particles TiO2 on Y-123 superconductor performance. J. Alloys Compd. 2019, 781, 664–673. [Google Scholar] [CrossRef]
  27. Chamekh, S.; Bouabellou, A. The Effects of Magnetic Dopant on the Structural and Electrical Properties in Superconducting YBaCu3O7-δ Ceramic. Adv. Chem. Eng. Sci. 2018, 8, 1–10. [Google Scholar] [CrossRef] [Green Version]
  28. Cava, R.J.; Batlogg, B.; Chen, C.H.; Rietman, E.A.; Zahurak, S.M. Oxygen stoichiometry, superconductivity and normal-state properties of YBa2 Cu3O7-o R. Nature 1987, 329, 9–11. [Google Scholar] [CrossRef]
  29. Benzi, P.; Bottizzo, E.; Rizzi, N. Oxygen determination from cell dimensions in YBCO superconductors. J. Cryst. Growth 2004, 269, 625–629. [Google Scholar] [CrossRef]
  30. Maaz, K.; Karim, S.; Mumtaz, A.; Hasanain, S.; Liu, J.; Duan, J. Synthesis and magnetic characterization of nickel ferrite nanoparticles prepared by co-precipitation route. J. Magn. Magn. Mater. 2009, 321, 1838–1842. [Google Scholar] [CrossRef]
  31. Hsiao, Y.-J.; Chang, Y.-H.; Chang, Y.-S.; Fang, T.-H.; Chai, Y.-L.; Chen, G.-J.; Huang, T.-W. Growth and characterization of NaNbO3 synthesized using reaction-sintering method. Mater. Sci. Eng. B Solid. State Mater. Adv. Technol. 2007, 136, 129–133. [Google Scholar] [CrossRef]
  32. Awano, M.; Fujishiro, Y.; Moon, J.; Takagi, H.; Rybchenko, S.; Bredikhin, S. Microstructure control of an oxide superconductor on interaction of pinning centers and growing crystal surface. Phys. C Supercond. Its Appl. 2000, 341–348, 2017–2018. [Google Scholar] [CrossRef]
  33. Uysal, E.; Ozturk, A.; Kutuk, S.; Çelebi, S. Effects of Lu Doping on the Magnetic Behavior of YBCO Superconductors Prepared by MPMG Method. J. Supercond. Nov. Magn. 2014, 27, 1997–2003. [Google Scholar] [CrossRef]
  34. Vanderbemden, P.; Cloots, R.; Ausloos, M.; Doyle, R.; Bradley, A.; Lo, W.; Cardwell, D.; Campbell, A. Intragranular and intergranular superconducting properties of bulk melt-textured YBCO. IEEE Trans. Appl. Supercond. 1999, 9, 2308–2311. [Google Scholar] [CrossRef]
  35. Arlina, A.; Halim, S.; Kechik, M.A.; Chen, S. Superconductivity in Bi–Pb–Sr–Ca–Cu–O ceramics with YBCO as additive. J. Alloys Compd. 2015, 645, 269–273. [Google Scholar] [CrossRef]
  36. Bean, C.P. Magnetization of High-Field Superconductors. Rev. Mod. Phys. 1964, 267, 31–39. [Google Scholar] [CrossRef]
  37. Clem, J.R. Granular and superconducting-glass properties of the high-temperature superconductors. Phys. C Supercond. Its Appl. 1988, 153–155, 50–55. [Google Scholar] [CrossRef]
  38. Ambegaokar, V.; Baratoff, A. Tunneling Between Superconductors. Phys. Rev. Lett. 1963, 11, 104. [Google Scholar] [CrossRef]
  39. Brecht, E.; Schmahl, W.; Miehe, G.; Rodewald, M.; Fuess, H.; Andersen, N.; Hanβmann, J.; Wolf, T. Thermal treatment of YBa2Cu3−xAlxO6+δ single crystals in different atmospheres and neutron-diffraction study of excess oxygen pinned by the Al substituents. Phys. C Supercond. Its Appl. 1996, 265, 53–66. [Google Scholar] [CrossRef]
  40. Presland, M.; Tallon, J.; Buckley, R.; Liu, R.; Flower, N. General trends in oxygen stoichiometry effects on Tc in Bi and Tl superconductors. Phys. C Supercond. Its Appl. 1991, 176, 95–105. [Google Scholar] [CrossRef]
  41. Bernhard, C.; Shaked, H. Generic superconducting phase behavior in high- Tc cuprates: Tc variation with hole concentration in YBa2Cu3O7-δ. Phys. Rev. B 1995, 51, 911–914. [Google Scholar]
  42. Panagopoulos, C.; Xiang, T. Relationship between the Superconducting Energy Gap and the Critical Temperature in High-Tc Superconductors. Phys. Rev. Lett. 1998, 81, 2336–2339. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) XRD patterns of Y123 samples with different wt.% of TiO2 nanoparticles; (b) XRD patterns of the peaks related to the (013) and (103) planes of Y123 with wt.% TiO2 nanoparticles.
Figure 1. (a) XRD patterns of Y123 samples with different wt.% of TiO2 nanoparticles; (b) XRD patterns of the peaks related to the (013) and (103) planes of Y123 with wt.% TiO2 nanoparticles.
Coatings 13 01093 g001
Figure 2. (af): FESEM images of surface morphology and EDX spectra along with histograms of average grain size distribution of Y123 with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.% of TiO2 addition.
Figure 2. (af): FESEM images of surface morphology and EDX spectra along with histograms of average grain size distribution of Y123 with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.% of TiO2 addition.
Coatings 13 01093 g002
Figure 3. AC susceptibility (χ = χ’ + iχ″) versus temperature of Y123/(TiO2)x at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%, measured in a magnetic field of 5 Oe and at a frequency of 295 Hz.
Figure 3. AC susceptibility (χ = χ’ + iχ″) versus temperature of Y123/(TiO2)x at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%, measured in a magnetic field of 5 Oe and at a frequency of 295 Hz.
Coatings 13 01093 g003
Figure 4. Derivative of the real part of AC susceptibility, dχ’/dT, as a function of temperature of Y123/(TiO2)x at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%.
Figure 4. Derivative of the real part of AC susceptibility, dχ’/dT, as a function of temperature of Y123/(TiO2)x at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%.
Coatings 13 01093 g004
Figure 5. (a) Normalised resistivity versus temperature curve of Y123 samples with various wt.% of TiO2; (b) derivative of resistivity against temperature of Y123 samples with various wt.% of TiO2.
Figure 5. (a) Normalised resistivity versus temperature curve of Y123 samples with various wt.% of TiO2; (b) derivative of resistivity against temperature of Y123 samples with various wt.% of TiO2.
Coatings 13 01093 g005
Figure 6. Comparison curves of Tc-onset measured from 4PP and ACS for Y123 at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.% TiO2 addition.
Figure 6. Comparison curves of Tc-onset measured from 4PP and ACS for Y123 at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.% TiO2 addition.
Coatings 13 01093 g006
Table 1. Lattice parameters a, b, and c, orthorhombicity, crystallite size, and oxygen content of Y123 with various wt.% of TiO2.
Table 1. Lattice parameters a, b, and c, orthorhombicity, crystallite size, and oxygen content of Y123 with various wt.% of TiO2.
TiO2 (x = wt.%)a (Å)b (Å)c (Å)Orthorhombicity (10−3)Crystallite Size (nm)Oxygen Content
0.03.8273.88511.6727.5201806.89
0.23.8283.88911.6957.90472.76.76
0.43.8263.88911.6838.16391.06.83
0.63.8243.88811.6908.29881.26.79
0.83.8293.89011.6917.90278.16.78
1.03.8253.88911.6958.29570.26.76
Table 2. EDX analysis of chemical composition of Y123 with x = (0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%) TiO2.
Table 2. EDX analysis of chemical composition of Y123 with x = (0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%) TiO2.
TiO2 Addition Atomic %
wt.%YBaCuOTiTotal
0.07.7614.7823.9653.510.0100%
0.27.3912.6625.1654.800.0100%
0.48.1017.4926.6747.740.0100%
0.66.4314.2422.3049.207.82100%
0.88.1817.3224.4050.100.0100%
1.03.1237.5112.4530.6816.25100%
Table 3. Summary data for Tc-onset, Tcj, Tp, Jc (Tpm), I0, and Ej of Y123/(TiO2)x at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%, measured in a magnetic field of 5 Oe and at a frequency of 295 Hz.
Table 3. Summary data for Tc-onset, Tcj, Tp, Jc (Tpm), I0, and Ej of Y123/(TiO2)x at x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0 wt.%, measured in a magnetic field of 5 Oe and at a frequency of 295 Hz.
TiO2 Addition
(x = wt.%)
Tc-onset (K)Tcj (K)Tp (K)Jc(Tpm) (A/cm2)I0 (μA)Ej × 10−21 J
TpgTpm
0.095.480.1-83.721.89.783.12
0.293.391.491.778.619.171.9223.01
0.492.891.391.778.520.890.1328.84
0.69391.391.88921.479.8725.55
0.892.690.390.4 7621.158.5318.72
1.093.291.591.981.121.580.2125.66
Table 4. Properties of superconducting transition temperature Tc-onset, Tc-offset, ΔTc, residual resistivity, ρ0, and hole concentration, p, of Y123 at various wt.% TiO2 addition.
Table 4. Properties of superconducting transition temperature Tc-onset, Tc-offset, ΔTc, residual resistivity, ρ0, and hole concentration, p, of Y123 at various wt.% TiO2 addition.
TiO2
Addition
Tc-onset (K)Tc-offset (K)Tc (K)ρ0 (Ω.cm)Hole Concentration, p
0.095.691.73.90.610.160
0.293.184.68.50.450.142
0.492.689.13.50.410.140
0.692.5893.50.440.140
0.891.287.240.370.138
1.094.791.730.300.149
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

Barood, F.; Kechik, M.M.A.; Tee, T.S.; Kien, C.S.; Pah, L.K.; Hong, K.J.; Shaari, A.H.; Baqiah, H.; Karim, M.K.A.; Shabdin, M.K.; et al. Orthorhombic YBa2Cu3O7−δ Superconductor with TiO2 Nanoparticle Addition: Crystal Structure, Electric Resistivity, and AC Susceptibility. Coatings 2023, 13, 1093. https://doi.org/10.3390/coatings13061093

AMA Style

Barood F, Kechik MMA, Tee TS, Kien CS, Pah LK, Hong KJ, Shaari AH, Baqiah H, Karim MKA, Shabdin MK, et al. Orthorhombic YBa2Cu3O7−δ Superconductor with TiO2 Nanoparticle Addition: Crystal Structure, Electric Resistivity, and AC Susceptibility. Coatings. 2023; 13(6):1093. https://doi.org/10.3390/coatings13061093

Chicago/Turabian Style

Barood, Fatma, Mohd Mustafa Awang Kechik, Tan Sin Tee, Chen Soo Kien, Lim Kean Pah, Kai Jeat Hong, Abdul Halim Shaari, Hussein Baqiah, Muhammad Khalis Abdul Karim, Muhammad Kashfi Shabdin, and et al. 2023. "Orthorhombic YBa2Cu3O7−δ Superconductor with TiO2 Nanoparticle Addition: Crystal Structure, Electric Resistivity, and AC Susceptibility" Coatings 13, no. 6: 1093. https://doi.org/10.3390/coatings13061093

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